KEMBAR78
Elwan Niversity F S: Aculty OF Cience P D | PDF | Electric Charge | Electric Field
0% found this document useful (0 votes)
15 views178 pages

Elwan Niversity F S: Aculty OF Cience P D

The document is a detailed outline of a physics course on Electricity and Magnetism at Helwan University, covering topics such as electrostatic charge, electric fields, Gauss's law, electric potential, capacitance, and magnetism. It includes sections on fundamental concepts, laws, and problems related to each topic, providing a comprehensive framework for understanding electromagnetism. The document is structured into chapters with specific subsections, illustrating key principles and experimental observations.

Uploaded by

01000621879ziad
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)
15 views178 pages

Elwan Niversity F S: Aculty OF Cience P D

The document is a detailed outline of a physics course on Electricity and Magnetism at Helwan University, covering topics such as electrostatic charge, electric fields, Gauss's law, electric potential, capacitance, and magnetism. It includes sections on fundamental concepts, laws, and problems related to each topic, providing a comprehensive framework for understanding electromagnetism. The document is structured into chapters with specific subsections, illustrating key principles and experimental observations.

Uploaded by

01000621879ziad
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/ 178

HELWAN UNIVERSITY

FACULTY OF SCIENCE
PHYSICS DEPARTMENT

0
10
40
51
50
30
0
10
Physics Department

40
51
Electricity and Magnetism
:‫إعداد‬
50
30

‫ أيمن مصيلحى‬.‫د‬

| P a g e II
Contents
1. Electrostatic Charge:
1.1. Introduction………………………………………........…1
1.2. Electric Charge .....…………………………........……….2

0
1.3. Conservation of charge……………………........…..……4

10
1.4. Quantization of Charge……………………........…..……5
1.5. Conductors and Insulators…………………..........….…...6

40
1.6. Coulomb Law………………………………........…...…..9
1.7. Multiple Forces….............................….……........…….11
51
Problems………………………………………….........….…15

2. Electrostatic Field:
50
2.1. Electric Field and Electric Forces……….........………..17
2.2. Electric Field Lines……………………….…………….21
30

2.3. Electric Charge in Electric Field…………….........…….24


2.4. Electric Dipole in Electric Field…………….........……..27
Problems………………………………………..............……33

3. Gauss Law:
3.1. Electric Flux………………………….............…………35
3.2. Gauss’s Law……………………….............……………41
Problems..……………………………….............…………...46
| P a g e III
4. Electric Potential:
4.1. Work…………………………….............………………47
4.2. Potential………………………….............……………...59
4.3. Equipotential Surfaces……………............…………….64

0
5. Capacitance:

10
5.1. Capacitor…………………………….............………….67
5.2. Capacitance………………………….............………….69
5.3. Electric-Field Energy……………….............…………..72

40
5.4. Dielectrics…………………………............……………75
51
6. Moving Charge:
6.1. Current……………………...............…………………...83
50
6.2. Resistance…………………...............…………………..89
30

7. Electromotive Force:
7.1. Circuits………………………………….............………97
7.2. Internal Resistance……………………...........………..101
7.3. Energy and Power in Electric Circuits.…...........……...104

| P a g e IV
8. Direct Current:
8.1. Kirchhoff’s Rules……………………...........…………107
8.2. Charging a Capacitore………………...........………….110
8.3. Discharging a Capacitor……………...........…………..117

0
9. Magnetism:

10
9.1. Introduction………………………...………………….121
9.2. Magnetic Field of Moving Charge……………...……..123

40
9.3. Magnetic Field of Current Element……………...…….128
9.4. Magnetic Field of Straight Conductor…………...……131
51
9.5. Magnetic Field of Circular Loop………………...…….136
50
10. Magnetic Field and Magnetic Force:
10.1.Magnetic Field………………...………….............…..141
30

10.2. Magnetic Field Line and Flux…………...........…..….148


10.3. Motion of Charged Particles in Magnetic Field….......154
10.4. Magnetic Force on Current-Carrying Conductor….....159
10.5. Force Between Parallel Conductors ……………........163
10.6. Ampere’s Law…………..…..……………………......167

References……………………………………………

|PageV
Chapter I: Electrostatic Charge

1.1. Introduction:
You are surrounded by devices that depend on the
physics of electromagnetism, which is the combination of

0
electric and magnetic phenomena such as computers,

10
television, radio, household lighting… etc. The physics of
electromagnetism was first studied by the early Greek

40
philosophers, who discovered that if a piece of, the straw will
jump to the amber (see Fig. 1.1). We now know that the
51
attraction between amber and straw is due to an electric force.
50
30

Figure (1.1): Electrified amber attracts feathers and confetti.

The Greek philosophers also discovered that if a certain


type of stone (a naturally occurring magnet) is brought near
bits of iron, the iron will jump to the stone. We now know that
the attraction between magnet and iron is due to a magnetic
2

force. The sciences of electricity and magnetism developed


separately for centuries—until 1820 when Hans Christian
Oersted found a connection between them: an electric current
in a wire can deflect a magnetic compass needle.

The new science of electromagnetism was developed

0
further by workers in many countries such as Michael Faraday.

10
In the mid-nineteenth century, James Clerk Maxwell put
Faraday’s ideas into mathematical form, introduced many new

40
ideas of his own, and put electromagnetism on a sound
theoretical basis. This chapter begins
51 with electrical
phenomena and discusses the nature of electric charge and
electric force.
50

1.2. Electric Charge:


30

A number of simple experiments demonstrate the


existence of electric forces. For example, after rubbing a
balloon on your hair on a dry day, you will find that the
balloon attracts bits of paper. The attractive force is often
strong enough to suspend the paper from the balloon. When
materials behave in this way, they are said to be electrified or
to have become electrically charged. In a series of simple
experiments, it was found that there are two kinds of electric
3

charges, which were given the names positive and negative by


Benjamin Franklin (1706–1790). Electrons are identified as
having negative charge, and protons are positively charged.

To verify that there are two types of charge, suppose a


glass rod is rubbed with a silk cloth and suspended by a string

0
to electrically isolate it from its surroundings as shown in Fig.

10
(1.2). When a second glass rod rubbed with silk cloth is
brought near the first one, the two rods repel each other. Next,

40
when we rub a plastic rod with fur and bring it near the
hanging glass rod, the rods are attracted to each other.
51
50
30

Figure (1.2) (a) Two charged rods of the same sign repel each other.
(b) Two charged rods of opposite signs attract each other.

This observation shows that the plastic and glass


have two different types of charge on them. Based on these
4

observations, we conclude that charges of the same sign repel


one another and charges with opposite signs attract one
another. Using the convention suggested by Franklin, the
electric charge on the glass rod is called positive and that on
the plastic rod is called negative. Therefore, any charged
object attracted to a charged plastic (rubber) rod (or repelled

0
by a charged glass rod) must have a positive charge, and any

10
charged object repelled by a charged plastic (rubber) rod (or
attracted to a charged glass rod) must have a negative charge.

40
The SI unit of charge is the coulomb. The coulomb unit
51
is derived from the SI unit ampere for electric current i. A
coulomb is defined as the amount of charge that flows through
50
a given cross-section of a wire in one second if there is a
steady current of 1 ampere in the wire:
30

Q=it
where "Q" is in coulombs if "i" in amperes and "t" is in
seconds. The ampere equals to one coulomb per second.

1.3. Conservation of Charge:


As deduced from experimental observations that electric
charge is always conserved in an isolated system. That is,
when one object is rubbed against another, charge is not
5

created in the process. The electrified state is due to a transfer


of charge from one object to the other. One object gains some
amount of negative charge while the other gains an equal
amount of positive charge. For example, when a glass rod is
rubbed on silk, the silk obtains a negative charge equal in
magnitude to the positive charge on the glass rod.

0
10
We now know from our understanding of atomic
structure that electrons are transferred in the rubbing process

40
from the glass to the silk. Similarly, when rubber is rubbed on
fur, electrons are transferred from the fur to the rubber, giving
51
the rubber a net negative charge and the fur a net positive
charge. This process works because neutral, uncharged matter
50
contains as many positive charges (protons within atomic
nuclei) as negative charges (electrons).
30

1.4. Quantization of Charge:


In 1909, Robert Millikan (1868–1953) discovered that
electric charge always occurs as integral multiples of a
fundamental amount of charge "e". In modern terms, the
electric charge q is said to be quantized, where q is the
standard symbol used for charge as a variable. That is, electric
charge exists as discrete “packets,” and we can write q= ± N e,
6

where N is some integer. Other experiments in the same period


showed that the electron has a charge -e while the proton has a
charge of equal magnitude but opposite sign + e. Some
particles, such as the neutron, have no charge.

1.5. Conductors and Insulators:

0
We can classify materials generally according to the

10
ability of charge to move through them. Conductors are
materials through which charge can move rather freely (such

40
as metals, human body, and tap water). Nonconductors—also
called insulators—are materials through which charge cannot
51
move freely; examples include rubber, plastic, glass, and
chemically pure water. Semiconductors are materials that are
50
intermediate between conductors and insulators; examples
include silicon and germanium. Superconductors are materials
30

that are perfect conductors, allowing charge to move without


any hindrance.

The properties of conductors and insulators are due to


the structure and electrical nature of atoms. Atoms consist of
positively charged protons, negatively charged electrons, and
electrically neutral neutrons. The protons and neutrons are
packed tightly together in the nucleus. An electrically neutral
7

atom contains equal numbers of electrons and protons. When


atoms of a conductor like copper come together to form the
solid, some of their outermost (most loosely held) electrons
become free to wander about within the solid, leaving behind
positively charged atoms (positive ions). We call the mobile
electrons conduction electrons. There are few (if any) free

0
electrons in a nonconductor.

10
Figure (1.3) demonstrates the mobility of

40
charge in a conductor. A negatively charged plastic rod will
attract either end of an isolated neutral copper rod. What
51
happens is that many of the conduction electrons in the closer
end of the copper rod are repelled by the negative charge on
50
the plastic rod. Some of the conduction electrons move to the
far end of the copper rod, leaving the near end depleted in
30

electrons and thus with an unbalanced positive charge. This


positive charge is attracted to the negative charge in
the plastic rod.
8

0
Figure (1.3): An isolated neutral copper rod attracted by a charged rod

10
Although the copper rod is still neutral, it is said to have

40
an induced charge, which means that some of its positive and
negative charges have been separated due to the presence of a
51
nearby charge. Similarly, if a positively charged glass rod is
brought near one end of a neutral copper rod, induced charge
50
is again set up in the neutral copper rod but now the near end
gains conduction electrons, becomes negatively charged, and
30

is attracted to the glass rod, while the far end is positively


charged.

Note that only conduction electrons, with their negative


charges, can move; positive ions are fixed in place. Thus, an
object becomes positively charged only through the removal of
negative charges.
9

1.6. Coulomb Law:


If two charged particles are brought near each other,
they exert an electrostatic force on each other. The direction of
the force vectors depends on the signs of the charges. If the
particles have the same sign of charge, they repel each other.
That means that the force vector on each is directly away from

0
the other particle (see Fig. (1.4)). If we release the particles,

10
they accelerate away from each other. If, instead, the particles
have opposite signs of charge, they attract each other. That

40
means that the force vector on each is directly toward the other
particle. If we release the particles, they accelerate toward each
51
other.
50
30

Figure (1.4): Two point charges separated by a distance "r" exert a force
on each other that is given by Coulomb’s law.
10

Charles Augustin Coulomb (1736 - 1806) studied the


inter-action forces of charged particles in detail in 1784. For
point charges (charged bodies that are very small in the
comparison with the distance r between them), Coulomb found
that the electric force is proportional to 1/r2. That is, when the
distance doubles, the force decreases to 1/4 of its initial value.

0
10
The force also depends on the quantity of charge, q, on
each body. He found that the forces that two point charges q 1

40
and q 2 a distance r apart exerts on each other is:
51
F=k
q1q 2
(1.1)
r2
50
where k is a proportionality constant called the electrostatic
constant or the Coulomb constant. Eqn. (1.1) is the
30

mathematical statement of Coulomb’s law: “The magnitude of


the force of interaction between two point charges is directly
proportional to the product of the charges and inversely
proportional to the square of the distance between them”.

The electrostatic constant k in Eqn. (1.1) is often written


as 1/4πε o . Then the magnitude of the electrostatic force in
Coulomb’s law becomes:
11

1 |q1 q2 |
F = 4πε (1.2)
0 r2

where ε 0 = 8.854 ×10-12 C2/N.m2


R R P P P P P P (or k = 8.99 ×109 P P

N.m2/C2).
P P P P

1.7. Multiple Forces:


U U

0
Electrostatic force obeys the principle of superposition.

10
Suppose we have "n" charged particles near a chosen
particle called particle 1; then the net force on particle 1 is

40
given by the vector sum:
𝐹⃗1,𝑛𝑒𝑡 = 𝐹⃗12 + 𝐹⃗13
51 + 𝐹⃗14 + ⋯ + 𝐹⃗1𝑛 (1.3)

𝐹⃗14 , for example, is the force on particle 1 due to the presence


of particle 4.
50

Example 1:
U
30

Three charged particles are


arranged in a line, as shown in Fig.
(1.5). Calculate the net
electrostatic force on particle 3
(the on the right) due to the other
two charges.
Figure (1.5) : Example 1
12

Solution:
The net force on particle 3 is the vector sum of the force
𝐹⃗31 exerted on particle 3 by particle 1 and the force 𝐹⃗32
exerted on particle 3 by particle 2:
𝐹⃗3 = 𝐹⃗31 +𝐹⃗32

0
• The magnitude of 𝐹⃗31 is:

10
|q1 q3 |
F31 = k
(𝑟31 )2

40
𝑚2 (8.00 × 10−6 𝐶)(4.00 × 10−6 𝐶)
9
= �8.988 × 10 𝑁. 2 �
𝐶 51(0.50 𝑚)2
= 1.2 𝑁
50
• The magnitude of 𝐹⃗32 is:
|q2 q3 |
30

F32 = k
(𝑟32 )2
𝑚2 (3.00 × 10−6 𝐶)(4.00 × 10−6 𝐶)
9
= �8.988 × 10 𝑁. 2 �
𝐶 (0.20 𝑚)2
= 2.7 𝑁
To get the direction of each force, let the line joining the
particles be the x axis, and we take it positive to the right.
Then, because 𝐹⃗31 is repulsive and 𝐹⃗32 is attractive, the
directions of the forces are as shown in Fig (1.5): F31 points in
13

the positive x direction (away from Q 1 ) and F32 points in the R R

negative x direction (toward Q 2 ). The net force on particle 3 is: R R

F 3 = - F 32 + F 31
R R R R R

= -2.7 N + 1.2 N = -1.5 N


The magnitude of the net force is 1.5 N, and it points to the
left.

0
10
Example 2:
U

Consider three point charges located at

40
the corners of a right triangle as shown
in Fig. (1.6), where q 1 = q 3 = 5.00 µC,51R R R R

q 2 = -2.00 µC, and a = 0.100 m. Find the


R R

resultant force exerted on q 3 . R R


50
Figure (1.6) : Example 2
Solution:
U
30

Charge q 3 will experience two electric forces exerted in


R R

different directions. The directions of the individual forces


exerted by q 1 and q 2 on q 3 are shown in Figure (1.6). The
R R R R R R

force 𝐹⃗32 exerted by q 2 on q 3 is attractive because q 2 and q 3


R R R R R R R R

have opposite signs. The direction of the attractive force 𝐹⃗23 is


to the left (in the negative x direction). The force 𝐹⃗31 exerted
14

by q 1 on q 3 is repulsive because both charges are positive. The


repulsive force 𝐹⃗31 makes an angle of 45.0º with the x axis.
• The magnitude of 𝐹⃗32 is:
|q2 q3 |
F32 = k
a2
𝑚2 (2.00 × 10−6 𝐶)(5.00 × 10−6 𝐶)
9
= �8.988 × 10 𝑁. 2 �

0
𝐶 (0.100 𝑚)2

10
= 8.99 𝑁
• The magnitude of 𝐹⃗31 is:

40
|q1 q3 |
F31 = k
51 (√2a)2
𝑚2 (5.00 × 10−6 𝐶)(5.00 × 10−6 𝐶)
9
= �8.988 × 10 𝑁. 2 �
𝐶 2(0.100 𝑚)2
50
= 11.2 𝑁
• The x and y components of the force 𝐹⃗31 :
30

F 31x = (11.2 N) cos 45.0º = 7.94 N


F 31y = (11.2 N) sin 45.0º = 7.94 N
• The net force acting on q 3 is:
F 3x = F 31x + F 32x = 7.94 N + (-8.99 N) = -1.04 N
F 3y = F 31y + F 32y = 7.94 N + 0 = 7.94 N
• Net force on q3 in unit-vector form:
𝐹⃗3 = (-1.04 𝚤̂ + 7.94 𝚥̂) N
15

Problems
U

1. Two point charges attract each other with an electric


force of magnitude F. If the charge on one of the particles is
reduced to one-third its original value and the
distance between the particles is doubled, what is the
resulting magnitude of the electric force between

0
them?

10
1 1 1 3 3
(a) F (b) F (c) F (d) F (e) F
12 3 6 4 2

40
2. The magnitude of the electric force between two 51
protons is 2.30 × 10-26 N. How far apart are they?
P P

(a) 0.1 m (b) 0.022 m (c) 3.1 m (d) 0.0057 m (e) 0.48 m
50
3. Three point charges lie along the
30

x axis as shown in Fig. (1.7). The


positive charge q 1 = 15.0 µC is at x
R R

= 2.00 m, the positive charge q 2 = R R

6.00 µC is at the origin, and the net


force acting on q 3 is zero. What is
R R

the x coordinate of q 3 ? R R 0T Figure (1.7): Problem 3


16

30
50
51
40
10
0
Chapter II: Electrostatic Field

2.1. Electric Field and Electric Forces:


Many common forces might be referred to as “contact
forces,” such as your hands pushing or pulling a cart. In

0
contrast, both the gravitational force and the electrical force

10
act over a distance. To understand the situation, the British
scientist Michael Faraday (1791–1867) introduced the idea of

40
the field. In the electrical case, according to Faraday, an
electric field extends outward from every charge and
51
permeates all of space (Fig. (2.1)). If a second charge is placed
near the first charge, it feels a force exerted by the electric
50
field that is there (say, at point P in Fig. (2.1)).The electric
field at point P is considered to interact directly with the
30

second charge producing force on it.

Figure (2.1): The electric field lines extending out from charge Q,
and P is an arbitrary point.
-18-

We can in principle investigate the electric field


surrounding a charge or group of charges by measuring the
force on a small positive test charge which is at rest. By a test
charge we mean a charge so small that the force it exerts does
not significantly affect the charges that create the field. If a
tiny positive test charge q is placed at various locations in the

0
vicinity of a single positive charge Q as shown in Fig. (2.2)

10
(points A, B, C), the force exerted on q is as shown. The
force at B is less than at A because B’s distance from Q is

40
greater (Coulomb’s law); and the force at C is smaller still. In
each case, the force on q is directed radially away from Q.
51
50
30

Figure (2.): The Force exerted by charge + Q on a small test charge, q,


0T 0T 0T

placed at points A, B, and C.

The electric field is defined in terms of the force on such


a positive test charge. In particular, the electric field, 𝐸�⃗ , at
any point in space is defined as the force 𝐹⃗ exerted on a tiny
-19-

positive test charge placed at that point divided by the


magnitude of the test charge q:
𝐹⃗
𝐸�⃗ = (2.1)
𝑞
More precisely, 𝐸�⃗ is defined as the limit of 𝐹⃗ /𝑞 as q
approaching zero. That is, q is so tiny that it exerts essentially

0
no force on the other charges which created the field. From

10
this definition (Eqn. (2.1)), we see that the electric field at any
point in space is a vector whose direction is the direction

40
of the force on a tiny positive test charge at that point, and
whose magnitude is the force per unit charge. Thus 𝐸�⃗ has SI
51
units of newtons per coulomb (N/C). The reason for defining
𝐸�⃗ as 𝐹⃗ /𝑞 (with q→0) is so that 𝐸�⃗ does not depend on the
50
magnitude of the test charge q. This means that 𝐸�⃗ describes
only the effect of the charges creating the electric field at that
30

point.

The electric field at any point in space can be measured,


based on the definition, Eqn.(2.1). For simple situations with
one or several point charges, we can calculate For example,
the electric field at a distance r from a single point charge Q
would have magnitude:
-20-

𝑘𝑞𝑄�
𝐹 𝑟2
𝐸= =
𝑞 𝑞
𝑄 1 𝑄
𝐸=𝑘 = (2.2)
𝑟2 4𝜋𝜀0 𝑟 2

Notice that E is independent of the test charge q—that

0
is, E depends only on the charge Q which produces the field,

10
and not on the value of the test charge q. Equation (2.2) is
referred to as the electric field form of Coulomb’s law.

40
If we are given the electric field 𝐸�⃗ at a given point in
51
space, then we can calculate the force 𝐹⃗ on any charge q
placed at that point by:
50

𝐹⃗ = 𝑞𝐸�⃗
30

(2.3)

If q is positive, 𝐹⃗ and 𝐸�⃗ point in the same direction. If q is


negative , 𝐹⃗ and 𝐸�⃗ point in opposite directions (see Fig.
(2.3)).
-21-

0
10
40
Figure (2.3): (a) Electric field at a given point in space. (b) Force on a
0T
51
positive charge at that point. (c) Force on a negative charge at that point.
50
2.2. Electric Field Lines:
U U

The electric field could be represented with arrows at


30

various points in a given situation, such as at A, B, and C in


Fig. (2.4). The directions of 𝐸�⃗𝐴 , 𝐸�⃗𝐵 , and 𝐸�⃗𝐶 are the same as for
the forces, but the magnitudes (arrow lengths) are different. To
indicate the electric field at many points, we draw a series of
lines to indicate the direction of the electric field at various
points in space. These electric field lines (or lines of force) are
drawn to indicate the direction of the force due to the given
field on a positive test charge.
-22-

0
10
0T Figure (2.4): Electric field vector, shown at three points, due to a single

40
point charge Q.

51
The lines of force due to a single isolated positive
charge are shown in Fig. (2.5a), and for a single isolated
50
negative charge in Fig. (2.5b). In part (a) the lines point
radially outward from the charge, and in part (b) they point
30

radially inward toward the charge because that is the direction


the force would be on a positive test charge in each case. We
can draw the lines so that the number of lines starting on a
positive charge, or ending on a negative charge, is proportional
to the magnitude of the charge. Notice that nearer the charge,
where the electric field is greater 𝐹 ∝ 1/r2, the lines are closer
P P

together. This is a general property of electric field lines: the


closer together the lines are, the stronger the electric field in
-23-

that region. Field lines can be drawn so that the number of


lines crossing unit area perpendicular to 𝐸�⃗ is proportional to
the magnitude of the electric field.

0
10
40
Figure (2.5): Electric field lines (a) near a single positive point charge,
0T 0T

510T

(b) near a single negative point charge.


50
Figure (2.6a) shows the electric field lines due to two
equal charges of opposite sign, a combination known as an
electric dipole. The electric field lines are curved in this case
30

and are directed from the positive charge to the negative


charge. The direction of the electric field at any point is
tangent to the field line at that point as shown by the vector
arrow 𝐸�⃗ at point P. Fig. (2.6b) shows the electric field lines
for two equal positive charges, and Fig. (2.6c) for unequal
charges, - Q and +2Q. Note that twice as many lines leave +2Q
-24-

as enter -Q (number of lines is proportional to magnitude of


Q).

0
10
40
51
50
30

Figure (2.6): Electric field lines for three arrangements of charges.


0T

2.3. Electric Charge in Electric Field:


U U

According to Eqn. (2.1), an electric field 𝐸�⃗ exerts a


force 𝐹⃗ on a point charge q given by:
𝐹⃗ = 𝐸�⃗ q. (2.4)
-25-

This force will produce an acceleration given as:


a =𝐹⃗ /m (2.5)
where m is the mass of the charged particle.

Consider two parallel metal plates, which are isolated from


each other, are connecting the terminals of a battery. If the

0
spacing between them is small compared with their

10
dimensions, the electric field between them will be fairly
uniform except near the edges. Two cases will be studied:

40
A) A charged particle of mass ‘m’ and charge ‘q’ is plased
51
at rest in this uniform field and then released as shown
in Fig.(2.7). The motion of the particle resembles that of
50
a body falling in the earth’s gravitational field.
According to Eqns.(2.4) and (2.5), the acceleration is
30

given by:
a = qE/m.

0T Figure (2.7): Charged particle in uniform electric field.


-26-

Apply the equations of uniformaly accelerated motion


with the initial velocity v o = 0, then,
R R

v = at
= qEt / m.
The displacement ‘y’ after time ‘t’ is given by,
y = ½ at2

0
P

= qEt2 / 2m,

10
P P

and
v2 = 2ay
P P

40
= 2qEy / m.
The kinetic energy K attained after moving this distance ‘y’ is
51
then,
K = ½ mv2 P
50
= ½ m (2qEy / m)
= qEy.
30

This is exactly the work due to a constant force (qE) acts a


distance (y).
B) A negative charged particle (electron) of mass ‘m’ and
charge ‘e’ is projected with speed v o at right angle to a R R

uniform electric field as shown in Fig.(2.8).The motion


resembles that of a projectile fired horizontally in the
earth’s gravitational field. The horizontal, x, and
vertical, y, motions are given by:
-27-

x = vo t
R R

and
y = ½ at2 P

= eEt2 / 2m. P P

Eliminating t yields,
y = eEx2 / 2mv o 2.

0
P P R RP P

10
40
51
Figure (2.8): Projected negative charge into uniform field.

Neglecting gravity, when the electron emerges from the


50
plates, it travels in a straigt line tangent to the parabola at the
exit point 0. Imagine a fluorescent screen S placed beyond the
30

plates, the electron will then make itself visible as a luminous


spot at P. This is exactly the principle of the electrostatic
cathode-ray tube.

2.4. Electric Dipole in Electric Field:


U U

An electric dipole is a pair of point charge with equal


magnitude and opposite sign, say, q and -q, separated by a
-28-

distance 1. In the following discussion, it is assumed that q is


positive. Two questions should be considered. First, what
forces and torques do an electric dipole experience when
placed in an external electric field (that is, a field set up by
charges outside the dipole)? Second, what an electric field
does on an electric dipole itself produce?

0
To start with the first question, let us place an electric

10
dipole in a uniform external electric field E (Fig. (2.9)). The
forces F+ and F- on the two charges both have magnitude qE,

40
but their directions are opposite and they add to zero. The net
force on an electric dipole in a uniform external electric field is
51
zero. However, the two forces do not act along the same line,
so their torque does not add to zero; tending to rotate the
50
dipole clockwise. They form a couple. Let the angle between
the electric field E and the dipole axis φ; then the torque τ
30

exerted by the couple is:


τ = Fr
= (qE)(1 sin φ) (2.6)
where r is the perpendicular distance between the lines of
action of the two forces = 1 sin φ.
-29-

Figure (2.9): Torque on electric dipole.

0
The product of the charge q and the separation 1 is the

10
magnitude of a quantity called “the electric dipole moment”,
denoted as p = q l. Its unit is charge times distance (C.m). Be

40
careful not to confuse it with momentum or pressure. There are
not as many letters in the alphabet as physical quantities; some
51
letters are used several times. Usually, the context makes it
clear what is meant.
50
In terms of p the torque τ exerted by the field has
magnitude
30

τ = p E sin φ (2.7)
This can be written more compactly in vector form. We define
the electric dipole moment to be a vector quantity p with
magnitude ql and the direction along the dipole axis from -ve
to +ve charge, then φ is the angle between the directions of the
vectors 𝑃�⃗and𝐸�⃗ , and:
𝜏⃗ = 𝑃�⃗x 𝐸�⃗ (2.8)
-30-

The torque is greatest when p and E are perpendicular


and is zero when they are parallel or antiparallel. The torque
always tends to turn p to line it up with E. The position φ = 0,
with p parallel to E, is a position of stable equilibrium, and
position φ = π/2, with p and E antiparallel is a position of
unstable equilibrium.

0
10
When a dipole changes direction in a field, the electric-
field torque does work on it, with a corresponding change in

40
potential energy. The work dW done by a torque τ during an
infinitesimal displacement dφ is given by;
dW = -τdφ.
51
The negative sign is found because the torque is in the
50
direction of decreasing φ. Then
dW = -τdφ
30

= - ρ E sinφ dφ
In a finite displacement from φ1 to φ2 the total work done on
R R R R

the dipole is:


φ2
W= ∫ − ρE sin φ dφ
φ1

= pE[cos φ]φφ 1

2
-31-

= ρEcosφ2 - ρEcosφ1
R R R

From the first law of thermodynamics, if dQ = 0 (adiabatic),


then
dW = -dU
Thus, the work is the negative of the change of potential
energy, W=U 1 -U 2 . So, a suitable definition of potential

0
R R R R

energy U for this system is:

10
U(φ) = -ρ E cosφ (2.9)

40
In Eqn. (2.9), we recognize the scalar product
𝜌⃗. 𝐸�⃗ = ρ E cosφ,
so we can also write:
51
U = - 𝜌⃗. 𝐸�⃗ (2.10)
50

The potential energy is minimum (most negative) at the


30

stable equilibrium position, where φ = 0 and p is parallel to E.


It is maximum (most positive) when φ = π and p is antiparallel
to E., at φ = π/2, where p is perpendicular to E, U is zero.
-32-

We summarize the properties of field lines as follows:


1. Electric field lines indicate the direction of the electric field;
the field points in the direction tangent to the field line at any
point.
2. The lines are drawn so that the magnitude of the electric
field, E, is proportional to the number of lines crossing unit

0
area perpendicular to the lines. The closer together the lines,

10
the stronger the field.
3. Electric field lines start on positive charges and end on

40
negative charges; and the number starting or ending is
proportional to the magnitude of the charge.
51
4. Electric fields also obey the principle of superposition. If
you want the net electric field at a given point due to several
50
particles, find the electric field due to each particle (such as 𝐸�⃗1 ,
due to particle 1) and then sum the fields as vectors:
30

𝐹⃗0 𝐹⃗01 𝐹⃗02 𝐹⃗03 𝐹⃗0𝑛


𝐸�⃗ = = + + +⋯+
𝑞0 𝑞0 𝑞0 𝑞0 𝑞0
𝐸�⃗ = 𝐸�⃗1 + 𝐸�⃗2 + 𝐸�⃗3 + ⋯ + 𝐸�⃗𝑛 (2.11)
-33-

Problems
U

1. Calculate the magnitude and direction of the electric field at


a point P which is 30 cm to the right of a point charge
Q = -3.0 × 10 -6 C.
P P

0
10
2. Two point charges are separated by a distance of 10.0 cm.
One has a charge of -25 µC and the other +50 µC. (a)

40
Determine the direction and magnitude of the electric field at a
point P between the two charges that is 2.0 cm from the
51
negative charge. (b) If an electron (mass = 9.11× 10-31 kg) is
P P

placed at rest at P and then released, what will be its initial


50
acceleration (direction and magnitude)?
30

0T Figure (2.10): Problem 2


-34-

3. Calculate the total electric field at point A in figure (2.11)


due to both charges, Q 1 and Q 2 .R R R R

0
10
40
0T Figure (2.11): Problem 3
51
4. Figure (2.12) shows three particles with charges q 1 =+2Q,R R
50
q 2 = -2Q, and q 3 = -4Q, each a distance d from the origin. What
R R R R

net electric field 𝐸�⃗ is produced at the origin?


30

0T Figure (2.12): Problem 4


Chapter III: Gauss Law

3.1. Electric Flux:

Consider an electric field that is uniform in both


magnitude and direction as shown in Fig. (3.1). The field lines

0
penetrate a rectangular surface of area whose plane is oriented

10
perpendicular to the field. The number of lines per unit area (in
other words, the line density) is proportional to the magnitude

40
of the electric field. Therefore, the total number of lines
penetrating the surface is proportional to the product EA. This
51
product of the magnitude of the electric field and surface area
perpendicular to the field is called the electric flux (Φ E ):
50
Φ𝐸 = 𝐸𝐴⊥ (3.1)
30

Figure (3.1): Field lines of a uniform


electric field penetrating a plane of
area perpendicular to the field.
-36-

From the SI units of E and A, we see that Φ E has units R R

of newton meters squared per coulomb (N.m2/C). If the surface


P P

under consideration is not perpendicular to the field, the flux


through it must be less than that given by Eqn. (3.1). Consider
Fig. (3.2a), where the normal to the surface of area A is at an
angle θ to the uniform electric field 𝐸�⃗ , the electric flux Φ E is

0
R R

defined as:

10
Φ𝐸 = 𝐸𝐴 cos 𝜃 (3.2)

40
where θ is the angle between the electric field direction and the
perpendicular to the area. 51
50
30

Figure (3.2): (a) A uniform electric field 𝐸�⃗ passing through a flat square
area A. (b)𝐸⊥ = 𝐸 cos 𝜃 is the component of 𝐸�⃗ perpendicular to the plane
of area A. (c) 𝐴⊥ = 𝐴 cos 𝜃 is the projection (dashed) of the area A
perpendicular to the field 𝐸�⃗ .
-37-

The flux can be written as:

Φ𝐸 = 𝐸𝐴⊥ = 𝐸⊥ 𝐴 = 𝐸𝐴 cos 𝜃 (3.3)

where 𝐸⊥ = 𝐸 cos 𝜃 is the component of 𝐸�⃗ perpendicular to


the area (Fig. (3.2b)) and, similarly, 𝐴⊥ = 𝐴 cos 𝜃 is the
projection of the area A perpendicular to the field 𝐸�⃗ (Fig.

0
(3.2c)).

10
According to Eqn. (3.3), we see that the flux through a
surface of fixed area A has a maximum value EA when the

40
surface is perpendicular to the field (when the normal to
the surface is parallel to the field, that is, when θ = 0º); the flux
51
is zero when the surface is parallel to the field (when the
normal to the surface is perpendicular to the field, that is, when
50
θ = 90º).

In more general situations, the electric field may vary


30

over a large surface. Therefore, the definition of flux given by


Eqn. (3.3) has meaning only for a small element of area over
which the field is approximately constant. Consider a general
surface of any shape that encloses a volume of space (such as
that shown in Fig. (3.3)), we divide the surface up into many
tiny areas ΔA 1 , ΔA 2 , ΔA 3 , …, and so on. We make the
R R R R R R

division so that each ΔA is small enough that it can be


-38-

considered flat and so that the electric field can be considered


constant through each ΔA. Then the total flux through the
entire surface is the sum over all the individual fluxes through
each of the tiny areas:

Φ E = E 1 ΔA 1 cos θ 1 + E 2 ΔA 2 cos θ 2 + ….
R R R R R R R R R R R R R R (3.4)

0
= � 𝐸𝑖 △ 𝐴𝑖 cos 𝜃𝑖 = � 𝐸�⃗𝑖 . Δ𝐴⃗𝑖

10
40
51
50
30

Figure (3.3): Electric field lines passing through a closed surface.

If the area of each element approaches zero, the number


of elements approaches infinity and the sum is replaced by an
integral. Therefore, the general definition of electric flux is:

Φ𝐸 = ∫ 𝐸�⃗ . 𝑑𝐴⃗ (3.5)


-39-

We are often interested in evaluating the flux through a


closed surface, defined as a surface that divides space into an
inside and an outside region so that one cannot move from one
region to the other without crossing the surface. The surface of
a sphere, for example, is a closed surface. By convention, if
the area element in Eqn. (3.5) is part of a closed surface, the

0
direction of the area vector is chosen so that the vector points

10
outward from the surface. If the area element is not part of a
closed surface, the direction of the area vector is chosen so that

40
the angle between the area vector and the electric field vector
is less than or equal to 90°.
51
Consider the closed surface in Fig. (3.4). The vectors
Δ𝐴⃗𝑖 point in different directions for the various surface
50
elements, but for each element they are normal to
the surface and point outward. At the element "1", the field
30

lines are crossing the surface from the inside to the outside and
θ < 90°; hence, the flux Φ E,1 =𝐸�⃗ . ∆𝐴⃗1 through this element is
R R

positive. For element 2, the field lines graze the surface


(perpendicular to Δ𝐴⃗2 ); therefore, θ= 90° and the flux is zero.
For elements 3, the field lines are crossing the surface from
outside to inside, 180 ˃ θ ˃ 90°and the flux is negative
because cos θ is negative.
-40-

0
10
40
51
Figure (3.4): A closed surface in an electric field.

The net flux through the surface is proportional to the


50
net number of lines leaving the surface, where the net number
means the number of lines leaving the surface minus the
30

number of lines entering the surface. If more lines are leaving


than entering, the net flux is positive. If more lines are entering
than leaving, the net flux is negative. The net flux Φ E through R R

a closed surface can be written as:

Φ𝐸 = ∮ 𝐸�⃗ ⋅ 𝑑𝐴⃗ = ∮ 𝐸𝑛 𝑑𝐴 (3.6)

where E n represents the component of the electric field normal


R R

to the surface.
-41-

3.2. Gauss's Law:


U U

An important relation in electricity is Gauss’s law,


developed by the great mathematician Karl Friedrich Gauss
(1777–1855). This relationship, describe a general relationship
between the net electric flux through a closed surface (often
called a Gaussian surface) and the charge enclosed by the

0
surface, and is a more general form of Coulomb’s law.

10
Consider a positive point charge q located at the center

40
of a sphere of radius r as shown in Fig. (3.5). The magnitude
of the electric field everywhere on the surface of the sphere is
E = k q/r 2. The field lines are directed radially outward and
P P
51
hence are perpendicular to the surface at every point on the
surface. That is, at each surface point, 𝐸�⃗ is parallel to the
50

vector ∆𝐴⃗𝑖 representing a local element of area ΔA i R R


30

surrounding the surface point. Therefore,

𝐸�⃗ ⋅△ 𝐴⃗𝑖 = 𝐸 △ 𝐴𝑖 (3.7)

and the net flux through the Gaussian surface is:

Φ𝐸 = ∮ 𝐸�⃗ ⋅ 𝑑𝐴⃗ = ∮ 𝐸 𝑑𝐴 = 𝐸 ∮ 𝑑𝐴 (3.8)

E is moved outside of the integral because, by symmetry, E is


constant over the surface.
-42-

0
10
Figure (3.5): A spherical gaussian surface of radius r surrounding a
positive point charge q.

40
The area of the spherical surface is ∮ 𝑑𝐴 = 𝐴 = 4𝜋𝑟 2 .
51
Hence, the net flux through the Gaussian surface is

𝑞 2)
1 𝑞
Φ𝐸 = 𝐸 � 𝑑𝐴 = 𝑘 2
(4𝜋𝑟 = 2
(4𝜋𝑟 2 )
𝑟 4𝜋𝜀0 𝑟
50
𝑞
Φ𝐸 = (3.9)
𝜀0
30

Equation (3.9) shows that the net flux through the


spherical surface is proportional to the charge inside the
surface and is independent of the radius. Now consider several
closed surfaces surrounding a charge q as shown in Fig. (3.6).
Surface S 1 is spherical, but surfaces S 2 and S 3 are not. From
R R R R R R

Eq. (3.9), the flux that passes through S 1 has the value q/ɛ 0 .
R R R R

As discussed before, flux is proportional to the number of


-43-

electric field lines passing through a surface. The construction


shown in Fig. (3.6) shows that the number of lines through S 1 R R

is
equal to the number of lines through the nonspherical surfaces
S 2 and S 3 . Therefore, the net flux through any closed surface
R R R R

surrounding a point charge q is given by q/ɛ 0 and is

0
R R

independent of the shape of that surface.

10
40
51
50
30

Figure (3.6): Closed surfaces of various shapes surrounding a charge q.

When many charges are enclosed by the surface,


Gauss's law can be written in the form:

𝜀0 Φ𝐸 = 𝑞𝑒𝑛𝑐

𝜀0 ∮ 𝐸�⃗ ⋅ 𝑑𝐴⃗ = 𝑞𝑒𝑛𝑐 (3.10)

where q enc is the net charges enclosed by the surface.


R R
-44-

In Eqn. (3.10), the net charge q enc is the algebraic sum


R R

of all the enclosed positive and negative charges, and it can be


positive, negative, or zero. If q enc is positive, the net flux is
R R

outward; if q enc is negative, the net flux is inward.


R R

Figure (3.7) shows two particles, with charges equal in


magnitude but opposite in sign, and the field lines describing

0
the electric fields the particles set up in the surrounding space.

10
Four Gaussian surfaces are also shown; let us consider each in
turn:

40
U Surface S 1 : The electric field is outward for all points on this
UR R

51
surface. Thus, the flux of the electric field through this surface
is positive, and so is the net charge within the surface (as
50
Gauss’ law requires).

Surface S 2 : The electric field is inward for all points on this


U UR R
30

surface. Thus, the flux of the electric field through this surface
is negative and so is the enclosed charge (Gauss’ law requires).

Surface S 3 : This surface encloses no charge, and thus q enc =


U UR R R R

0. Gauss’ law (Eqn. (3.10)) requires that the net flux of the
electric field through this surface be zero. That is reasonable
because all the field lines pass entirely through the surface,
entering it at the top and leaving at the bottom.
-45-

Surface S 4 : This surface encloses no net charge, because the


U UR R

enclosed positive and negative charges have equal magnitudes.


Gauss’ law requires that the net flux of the electric field
through this surface be zero. That is reasonable because there
are as many field lines leaving surface S 4 as entering it.
R R

0
10
40
51
50
30

Figure (3.7): Two charges, equal in magnitude but opposite in sign,


and the field lines that represent their net electric field.
-46-

Problems
U

1. A thin spherical shell of radius r0 possesses a total net


charge Q that is uniformly distributed on it, Fig. (3.8).
Determine the electric field at points (a) outside the shell, and

0
(b) inside the shell.

10
40
Figure (3.8): Problem 1. 51
50
2. A spherical gaussian surface surrounds a point charge q.
30

Describe what happens to the total flux through the surface


if (A) the charge is tripled, (B) the radius of the sphere is
doubled, (C) the surface is changed to a cube, and (D) the
charge is moved to another location inside the surface.

3. A point charge 𝑞 = 3.0 𝑛𝐶 is at the center of a cube with


sides of length 0.20 m. what is the electric flux through one
of the six faces of the cube?
Chapter IV: Electric Potential

4.1. Work:
When a force F acts on a particle that moves from point
‘a’ to point ‘b’, the work Wa → b done by the force is

0
b b
Wa→b = ∫ F .dl = ∫ F cos φdl (4.1)

10
a a

where dl is an element of the path of the particle and φ is the

40
angle between F and dl at each point along the path.
The force field is conservative, this work can always be
51
expressed in terms of a potential energy U. When the particle
moves from a point where this potential energy is U a , to a
50
point where it is U b , the work Wa → b done by the force (field)
on the particle is
30

Wa→b = U a − U b (4.2)
where Wa → b is positive, U a is greater than U b , and the
potential energy decreases. That is what happens when a
baseball falls from a high point under the action of the earth’s
gravity. The force of gravity does positive work, and the
gravitational potential energy decreases. When a ball is thrown
upward, the gravitational force does negative work during the
ascent, and the potential energy increases.
-48-

The work-energy theorem says that the change in kinetic


energy (k a -k b ) during any displacement is equal to the total
R R R R

work done on the particle. So, if Eqn.(4-2) gives the total


work, then
k a - k b = U b -U a ,
R R R R R R R R

or

0
k a + U a = k b +U b (4.3)

10
R R R R R R R R

A) Point charge in a uniform electric field:

40
In Fig.(4.1) a pair of charged parallel metal plates
sets up a uniform electric field with magnitude E. The
51
field exerts a downward force with magnitude F = q`E on
a positive test charge q` as the charge moves a distance d
50
from point a to point b. The force of the test charge is
constant, independent of its location, so the work done by
30

the electric field is


Wa → b = F . d = q' Ed (4.4)
This work can be represented with potential-energy
function U. The y-component of force F = -q` E is constant,
and there is no x- or z-component, so the work is independent
of the path the particle takes from ‘a’ to ‘b’. Just as the
-49-

potential energy for the gravetional force F = -mg was, U = m


g y, the potential energy for the electric field force F= -q` E is

U = q' Ey (4.5)

0
10
40
51
50

Figure (4.1): Test charge moving in uniform electric field.


30

When the test charge moves from height ‘a’ too high
‘b’, the work done on the charge by the field is given by:
Wa → b = U a − U b
= q ` E y a − q `E y b (4-6)
= q` E( y a − y b )
-50-

When y a is greater than y b (Fig. 4.2a), the particle


R R R R

moves in the same direction as the E field, U decreases, and


the field does positive work. When y a is less than y b (Fig.
R R R R

4.2b), the particle moves in the opposite direction to E, U


increases, and the field does negative work. In particular, if y a R R

=d and y b = 0, then Eqn. (4.6) gives W a→b = q’E d, in

0
R R R R

agreement with Eqn. (4.4).

10
40
51
50
30

Figure (4.2): Positive charge moves in: a) direction of the electric field,
and b) direction opposite to the electric field.

If the test charge q’ is negative, the potential energy


increases when it moves with the field and decreases when it
moves against the field Fig. (4.3).
-51-

0
10
Figure (4.3): Negative charge moves in: a) direction of the electric

40
field, and b) direction opposite to the electric field.

Any charge distribution can be represented as a collection of


51
point charges. Therefore, it is useful to calculate the work
done on a test charge q` moving in the electric field caused by
50
a single stationary point charge q. First, a displacement along
the radial line will be considered as shown in Fig. (4.4), from
30

point a to point b. The force is not constant, and we have to


integrate to calculate the work done on q`. The force on q` is
given by Coulomb law, and its radial component is:
1 qq′
Fr = (4.7)
4πε o r 2
R R
-52-

0
10
40
Figure (4.4): Test charge moves along radial straight line from q.
51
If q or q’ is negative, F is also negative; if both q and q’
are negative, F is positive. The work W a→b done on q’ by force
R R
50
F as q’ moves from a to b is:
rb


Wa→b = Fr dr
30

ra
rb
1 qq\
= ∫ 4πε o r 2
dr
ra
b
q q \  1
= −
4πε o  r  a
qq ′ 1 1
= ( − ) (4.8)
4πε o ra rb
-53-

Thus, the work for this particular path depends only on


the end points. In fact, the work is the same for all possible
paths from a to b. To prove this, a more general displacement
is considered (Fig.4.5) in which ‘a’ and ‘b’ do not lie on the
same radial line. The work done on q` during this
displacement is given by:

0
rb
W a→b = ∫ F cosφ dl

10
R R (4.9)
ra

40
51
50
30

Figure (4.5): The work done on charge q 0 by the electric field of charge q
R R

does not depend on the path taken, but only on the distances r a and r b .
R R R R

However, the figure shows that cos φ dl = dr. That is,


the work done during a small displacement dl depends only on
-54-

the change dr in the distance r between the charges, which is


the radial component of the displacement. Thus, Eqn. (4.9)
gives the work done by the field even for this more general
displacement.
Equation (4.9) shows that the work done on q’ by the E
field produced by q depends only on r a and r b , not on the

0
R R R R

details of the path. Also, if q’ returns to its starting point a by

10
a different path, the total work done in the round-trip
displacement is zero. These are the needed characteristics for

40
a conservative force field. Thus the force on q’ is a
conservative force field. 51
B) Potential energy:
50
Comparing Eqns. (4.2) and (4.8), they are consistent if
we define qq ' / 4πε o ra to be the potential energy U a when q` is
R R
30

at point ‘a’, a distance r a from q, and qq ' / 4πε o rb to be the


R R

potential energy U b when q` is at point ‘b’, a distance r b from


R R R R

q. Thus, the potential energy U when the test charge q` is at


any distance r from charge q is:
qq′
U= (4.10)
4πε o r
-55-

Note that, there is nothing assumed about the signs of q and q`,
therefore, Eqn. (4.10) is valid for any combination of signs.
Gauss’s law tells us that the electric field outside any
spherically symmetric charge distribution is the same as
though all the charge were concentrated at the center.
Therefore, Eqn. (4.10) also holds if the test charge q` is outside

0
any spherically symmetric charge distribution with total charge

10
q, at a distance r from the center.
Potential energy is always defined relative to some

40
reference point where U=0. In Eqn. (4.10), U is zero when q
and q` are infinitely far apart, or r = ∞.
51 Therefore, U
represents the work done on the test charge q’ by the field of q
when q` moves from an initial distance r to infinity. If q and
50
q` have the same sign, the interaction is repulsive, this work is
positive, and U is positive at any finite separation. If they
30

have opposite sign, the interaction is attractive and U is


negative.
Equation (4.10) can be generalized for situation that
the E field, in which charge q` moves, is caused by several
point charges q 1 , q 2 , q 3 , ... at distances r 1 , r 2 , r 3 , ... from q`.
R R R R R R R R R R R R

The total electric field at each point is the vector sum of the
field due to the individual charges, and the total work done on
-56-

q` during any displacement is the sum of the contributions


from the individual charges.

The potential energy associated with a test charge q` at


point ‘a’ in Fig.(4.6), due to a collection of charges q 1 , q 2 , q 3 , R R R R R R

... at distances r 1 , r 2 , r 3 , ... from the test charge q`, is the

0
R R R R R R

algebraic sum (not a vector sum), is then given as:

10
q ` q1 q 2 q 3
U= ( + + + ...)
4πε o r1 r2 r3

40
q` qi
= ∑
4πε o i ri 51 (4.11)
50
30

Figure (4.6): Potential energy of collection of charges.

When q` is at a different point ‘b’, the potential energy


is given by the same expression, but r 1 , r 2 , ... are the distances
R R R R

from q 1 , q 2 , ... to the point ‘b’. The work done on charge q`


R R R R
-57-

when it moves from ‘a’ to ‘b’ along any path is equal to the
difference between the potential energies U a and U b when q` is R R R R

at a and b, respectively. Thus,


Wa → b = U a - U b
R R R R R R R

Any charge distribution can be represented as a

0
collection of point charges, so Eqn. (3.11) shows that a

10
potential-energy function can be always found for any static
electric field. It follows that every electric field due to a static

40
charge distribution is a conservative force field.
51
Equations (4.10) and (4.11) define U to zero when all
the distances r 1 , r 2 , ... are infinite, that is, when the test charge
R R R R
50
q` is very far away from all the charges that produce the field.
This position is in a sense equidistant from all the charges q
30

(except when the charge distribution itself extends to infinity).


As with any potential-energy function, the reference point is
arbitrary; we can always add a constant to make U equal zero
at any point we choose. Making U=0 at infinity is a
convenient eference level for electrostatic problems, but in
circuit analysis other reference levels are often more
convenient.
-58-

Equation (4.11) gives the potential energy associated


with the presence of the test charge q` in the E field produced
by q 1 , q 2 , q 3 , ... . However, there is potential energy involved
R R R R R R

in assembling these charges. If we start with charges q 1 , q 2 , R R R R

q 3 , ... all separated from each other by infinite distances and


R R

then bring them together so that the distance between q 1 and q`

0
R R P P

is r, the total potential energy is the sum of the pair

10
interactions. It can be written as:
1 qi q j

40
U= (4.12)
4πε o ii rij

This sum extends over all pairs of charges; counting


51
each pair only once and excluding the interaction of a charge
with itself (i = j). If a term with i=3 and j=4 is included, a term
50
with j = 3 and i = 4 should not be included. A neat way to
express this limitation is to sum only over i’s and j’s for which
30

i > j.
Here is a final comment about electric potential energy.
It has been defined in terms of the work done by the electric-
field force on a charged particle moving in the field. When a
particle moves from point ‘a’ to point ‘b’, the work done on it
by the electric field is:
Wa→b = U a -U b .
R R R R R R
-59-

When U a is greater than U b , the field does a positive work on


R R R R

the particle as it falls from a point ‘a’ of higher potential


energy to a point ‘b’ of lower potential energy. An alternative
but equivalent viewpoint is that: in order to raise a particle
from appoint ‘b’ where the potential energy is U b to a point ‘a’
R R

of a greater potential energy U a , it would be need to apply an

0
R R

additional force F, which is equal and opposite of the electric-

10
field force and does positive work. This is equivalent to
pushing two similar charges closer together. The potential-

40
energy difference (U a -U b ) is then defined as the work done by
R R R R

F ext during the reverse displacement from ‘b’ to ‘a’ because


R R
51
F ext is the negative of the electric-field force and the
R R

displacement is in the opposite direction. This alternative


50
viewpoint will not be used because the hypothetical additional
force E ext may not actually present in a particular problem.
R R
30

4.2. Potential:
U U

In the first section we looked at the potential energy U


associated with a test charge q` in an electric field. Now, this
potential energy will be described on a “per unit charge” basis;
leading to the concept of electric potential, often called simply
potential. It is the potential “energy per unit charge”. This is
just as electric field describes the force on a charge particle in
-60-

the field on a “force per unit charge” basis. This concept is


very useful in calculations involving energies of charged
particles. It also facilitates many electric-field calculations
because it is closely related to the E field. When we need to
calculate an electric field, it is often easier to calculate the
potential first and then find the field from it.

0
The potential V at any point in an electric field is

10
defined as the potential energy U per unit charge associated
with a test charge q` at that point:

40
U
V=
q′ 51
or
U = q′ V (4.13)
50
Potential energy and charge are both scalars, so
potential is a scalar quantity. From Eqn. (4.13), its unit is
30

energy divided by charge. The SI unit of potential, J/C, is


called volt (V) in honor of the scientist Alexandro Volta
(1745-1827).
1 V = 1 J/C = 1 joule/coulomb.
In circuit analysis, potential is often called voltage. To
put Eqn.(4.2) on a “work per unit charge” basis, both sides are
divided by q`, obtaining :
-61-

Wa→b U a U b
= − = Va − Vb (4.14)
q′ q′ q′

where V a = U/q` is the potential energy per unit charge at


R R

point ‘a’ and V b is that at ‘b’. V a and V b are called the


R R R R R R

potential at point ‘a’ and potential at point ‘b’, respectively.

0
To find the potential V at a point due to any collection of point

10
charges, Eqn. (4.11) is divided by q`:

40
U 1 q
V= =
q′ 4πε o
∑ ri (4.15)
i i 51
When we have a continuous distribution of charge along
a line, over a surface, or through a volume, we divide the
50
charge into elements dq and the sum becomes an integral:
1 dq
V= ∫ (4.16)
30

4πε o r
where r is the distance from the charge element dq to the field
point at which V is defined.
In deriving Eqn. (4.16), an expression is has been used
for the potential of a point charge that is zero at an infinite
distance from the charge, thus, the V defined by Eqn. (4.16) is
zero at points infinitely far away from all the charges. Later we
shall encounter cases in which the charge distribution itself
-62-

extends to infinity, and we won’t be able to use this equation


directly.
When we are given a collection of point charges,
Eqn.(4.16) is usually the easiest way to calculate the potential
V. However, in some problems in which the E field is known
or can be found easily, it is easier to work directly with the

0
field. The force F on the test charge q` can be written as:

10
F = qE,
so

40
b
Wa → b = ∫ F.dl
a 51
b
= ∫ q ′Edl.
a
50
When this is combined with Eqn. (4.14), the test charge
q`, cancels, and one gets:
30

b
Va − Vb = ∫ E.dl
a
(4.17)
b
= ∫ E cos φdl.
a

When the E field does positive work on a positive test


charge that moves from a to b, the potential must be greater at
a than at b. Or, using the alternative viewpoint mentioned at
the end of section 3-1, we say that to move a test charge from
-63-

b to a would require an external force per unit charge to -E.


The work done by this external force is the potential difference
(V a - V b ). That is :
R R R R

a
Va − Vb = − ∫ Edl
b
a
= − ∫ F.dl / q'

0
b

10
Compared to Eqn. (4.17), this has negative sign and the limits
are reversed. Therefore the two expressions are equal.

40
The difference (V a - V b ) is called the potential of ‘a’
R
51 R R R

with respect to ‘b’; this difference is sometimes abbreviated as:


V ab = V a - V b .
R R R R R R
50
This is often called the potential difference between ‘a’ and ‘b’
but that is ambiguous unless the reference point is specified. In
30

addition to the moving coil voltmeter, there are also much


more sensitive potential-measuring devices that use electronic
amplification. Instruments that can measure a potential
difference of 1µV are common, and sensitivities down to 10- P

12
V can be attained.
P

Note that potential, like electric field, is independent of


the test charge q, that we use to define it. When a positive test
-64-

charge moves from higher to lower potential (that is V a >V b ), R R R R

the electric field does positive work on it. A positive charge


tends to “fall” from a high potential region to a lower potential
one. The opposite is true for negative charge.
Potential calculations usually follow one of two routes.
If we know the charge distribution, we can use Eqn. (4.15) or

0
(4.16). Or if we know the electric field, we can use Eqn.

10
(4.17), defining the potential to be zero at same convenient
place. Some problems require a combination of these

40
approaches.
1- Parallel Plates. 51
2- Line charge and charged conducting cylinder.
3- Charged circular ring.
50
4- Charged thin rod.
30

4.3. Equipotential Surfaces:


U U

Field lines help us visualize electric fields. In a similar


way the potential at various points in an electric field can be
represented graphically by equipotential surfaces. An
equipotential surface defined as a surface on which the
potential is the same at every point. In region where an
electric field is present, we can construct an equipotential
surface through any point. Figure (4.7) shows a some
-65-

equipotential surfaces, often with equal potential differences


between adjacent ones, and electric field lines for assemblies
of point charges. No point can be at two different potentials, so
equipotential surfaces for different potentials can never touch
or intersect.

0
10
40
51
50
30

Figure (4.7): Equipotential surfaces and electric field lines for: a) single
isolated positive charge, b) electric dipole and c) two equal positive
charges.

The potential energy for a test charge is the same at


every point on a given equipotential surface, so the E field
does no work on a test charge when it moves from point to
-66-

point on such a surface. It follows that the E field can never


have a component tangent to the surface; such a component
would do work on a charge moving on the surface. Therefore
E must be perpendicular to the surface at every point.
Field lines and Equipotential surfaces are always
mutually perpendicular. In general, field lines are curves and

0
equipotentials are curved surfaces. For the special case of a

10
uniform field, in which the field lines are straight, parallel, and
equally spaced, the equipotentials are parallel planes

40
perpendicular to the field lines.
An arbitrary family of equipotential surfaces are shown
51
in Fig.(4.8). Science paths I and II begin and end on the same
equpotential surface, the work to move a charge along these
50
paths is zero. On the other hand, the work to move a charge
along paths I` and II` is not zero. However, science the two
P P P P
30

paths connect the same pair of equipotential surfaces, the work


is the same for each path because the initial and the final
potentials are the same.

Figure (4.8): Different


paths of a test charge
along different
equipotential surfaces.
Chapter V: Capacitance

5.1. Capacitor:
Any two conductors separated by an insulator form a

0
capacitor (Fig. 5.1). In most practical applications the

10
conductors have charges with equal magnitude and opposite
sign, and the net charge on the capacitor as a whole is zero.

40
When we say that a capacitor has charge Q, we mean that the
conductor at higher potential has charge +Q and the conductor
51
at lower potential has charge - Q.
50
30

Figure (5.1): Two different insulated charges.


-68-

The electric field at any region between the conductors


is proportional to the magnitude Q of charge on each
conductor. It follows that the potential difference between the
conductors is V, that is, the potential of the positively charge
conductor with respect to the negatively charge conductor is
also proportional to Q. If the magnitude of charge on each

0
conductor is doubled, the charge density at each point doubles,

10
and the potential difference between conductors doubles but
the ratio of charge to potential difference does not change.

40
The capacitance C of a capacitor is defined as the ratio
51
of the magnitude of the charge Q on either conductor to the
magnitude of the potential difference V ab between the
R R
50
conductors :
Q
C= (5.1)
30

Vab
The SI unit of capacitance is called one Farad (1F), in
honor of Micheal Faraday. From Eqn. (5.1), Farad is equal to
one coulomb per volt:
1F = 1 C/V
Capacitors have thousands of practical uses, and
contemporary electronics could not exist without them.
Capacitors are used in energy-storage units for pulsed lasers,
-69-

in computer chips that store information, in circuits that


improve the efficiency of ac power transmission lines, and in
many other areas. The study of capacitors will also help us to
develop insight into the behaviour of electric fields and their
interactions with matter.

0
5.2. Capacitance:

10
U U

The most common form of capacitor consists of two


parallel conducting plates, each with area A separated by a

40
distance ‘d’ that is small in comparison with there dimensions
(Fig.5.2). 51
50
30

Figure (5.2): Parallel-plate capacitor.

Nearly all the field of such a capacitor is localized in the


region between the plates, as shown. There is some “fringing”
of the field at the edges, as shown in the figure, but if the
distance between plates is small in comparison to their size, we
-70-

can neglect this. The field between the plates is then uniform
and the charge on the plates are uniformly distributed over
their opposing surfaces. This arrangement is called a parallel-
plate capacitor. The electric-field magnitude is given as
E=σ/ε o ,
R R

where σ is the magnitude of surface charge density (charge per

0
unit area) on each plate. This is equal to the magnitude of total

10
charge Q on each plate divided by the area A of the plate, or
σ = Q/A,

40
so the electric field E can be expressed as:
σ Q
E= =
51 εo εo A
The field is uniform; if the distance between the plates is
50
d, then the potential difference (voltage) between them is:
V∞ = Ed
30

Qd
=
εo A

From this it can be seen that the capacitance C of a


parallel-plate capacitor in vacuum is:
Q
C=
Vab
(5.2)
A
= εo
d
-71-

The quantities ε o , A, and d are constants for a given


R R

capacitor. The capacitance C is, therefore, a constant which is


independent of the charge on the capacitor. It is directly
proportional to the area A of each plate and inversely
proportional to their separation d.
In Eqn. (5.2), if A is in square meters and d in meters, C

0
is in farads. The units of ε o are C2/N.m2, so that:

10
R R P P P P

1F = 1C2/N.m P P

= 1C2/J

40
P P

Because 1 V = 1 J/C (energy per unit charge), this is


consistent with our definition F = C/V. Finally, the units of
51
ε o can be expressed as:
R R

1C2/N.m2 = 1F/m
50
P P P P

This relation is useful in capacitance calculations, and it


also helps us to verify that Eqn. (5.2) is dimensionally
30

consistent.
It has been assumed that there is only vacuum in the
space between the plates. When matter is present, things are
somewhat different. Meanwhile, we remark that if the space
contains air at atmospheric pressure instead of vacuum, the
capacitance differs from prediction of Eqn. (5.2) by less than
0.06%.
-72-

In many applications the most convenient units of


capacitance are the microfarad (1µf=10-6 F) and the picofarad
P P

(1pF = 10-12 F). For example, the power supply of an ac-


P P

power AM radio contains several capacitors with capacitances


of the order of 10 or more microfarads, and the capacitances in
the tuning circuits are of the order of 10 to 100 picofarads.

0
10
5.3. Electric-Field Energy:
U U

Many of most important applications of capacitors

40
depend on their ability to store energy. The opposite charges
on the plates, separated and attracted toward each other are
51
analogous to a stretched spring or a body lifted in the earth’s
gravitational field. The potential energy corresponds to the
50
energy input required to charge the capacitor and to the work
done by the electrical forces when it becomes discharged. This
30

work is analogous to the work done by a spring or the earth’s


gravity when the system returns from displaced position to the
reference position.
One way to calculate the potential energy U of a
charged capacitor to calculate the work W required to charge
it. The final charge Q and the final potential difference V are
related by :
Q = CV
-73-

Let V and q be the varying potential difference and


charge during the charging process: then V = q/C. The work
dW required to transfer an additional element of charge dq is:
dW = V dq
qdq
=
C

0
The total work W needed to increase the charge q from

10
zero to a final value Q is:
w
W = ∫ dW

40
o

1Q Q2
=51 ∫
Co
q dq =
2C
This is also equal to the total work done by the electric
50
field on the charge when q decreases from an initial value Q to
zero, as the elements of charge dq “fall” through potential
30

differences v that vary from V down to zero.


If we define the potential energy of an uncharged
capacitor to be zero, then W is equal to the potential energy U
of the charged capacitor. The final potential difference V
between the plates is:
V = Q/C,
So, U (which is equal to W) can be expressed as:
-74-

Q2
U=
2C
1
= QV (5.3)
2
1
= CV 2
2
When Q is in coulombs, C in farads (coulombs per volt), and

0
V in volts (joules per coulomb), U is in joules.

10
The last form of Eqn. (5.3) also shows that the total

40
work W is equal to the total charge Q transferred, times the
average potential difference V/2 during charging.
51
A charged capacitor is the electrical analog of a
50
stretched spring with elastic potential energy U = ½ kx2. The
P P

charge Q is analogous to the elongation x, and the reciprocal


30

of the capacitance, 1/C, is analogous to the force constant k.


The energy supplied to a capacitor in the charging process is
analogous the work do on a spring when it is stretched.
The energy stored in a capacitor is related directly to the
electric field between the capacitor plates. In fact, we can think
of the energy as stored in the field in the region between the
plates. To develop this relation, let us find the energy per unit
volume in the space between the plates of a parallel-plate
-75-

capacitor with plate area A and separation d. We call this the


energy density, demoted by u. From Eqn. (5.3) the total energy
is ½ CV2, and the volume between the plates is simply Ad.
P P

The energy density is:


1 / 2 (CV 2 )
u=
Ad

0
10
The potential difference is related to the electric field
magnitude E by V = Ed. Then, using Eqn. (5.2), one gets:

40
1
u= εoE2 (5.4)
51 2
Although this relation has been derived only for one
specific kind of capacitor, it turns out to be valid for any
50
capacitor in vacuum and indeed for any electric field
configuration in vacuum. This result has an interesting
30

implication. We think of vacuum as space with no matter in it,


but vacuum can nevertheless have electric fields and therefore
energy. It is not necessarily just empty space.

5.4. Dielectrics:
U U

Most capacitors have none conducting material, or


dielectric, between their plates. A common type of capacitor
uses long strips of metal foil for the plates, separated by strips
-76-

of plastic sheet. A sandwich of these materials is rolled up,


forming a unit that can provide a capacitance of several
microfarads in a compact package.
Electrolytic capacitors use as their dielectric an
extremely thin layer of non-conducting oxide between a metal
plate and conducting solution. From Eqn. (5.2) the capacitance

0
is inversely proportional to the distance d between the plates.

10
Because of the thinness of the dielectric, electrolytic capacitors
with relatively small dimensions may have a capacitance of the

40
order of 100 or 1000 µF.
Placing a solid dielectric between the plates of a
51
capacitor serves three functions. First, it solves the mechanical
problem of maintaining two large metal sheets at a very small
50
separation without actual contact.
30

Second any dielectric material, when subjected to a


sufficiently large electric field, experiences dielectric
breakdown a partial ionization that permits conduction through
a material that is supposed to insulate. Many insulating
materials can tolerate stronger electric fields without
breakdown than can air.
-77-

Third, the capacitance of a capacitor of given


dimensions becomes greater when there is a dielectric material
between the plates than with air or vacuum. We can
demonstrate this effect with the aid of a sensitive electrometer.
A charged capacitor, with charge of magnitude Q on
each plate and potential difference V o . When we insert a sheet

0
R R

of dielectric such as glass, paraffin, or polystyrene, between

10
the plates, the potential difference decrease to a smaller value
v. When we remove the dielectric, the potential difference

40
returns to its original value V o , showing that the original
R R

charges on the plates have not changed. 51


The original capacitance C O is given by C O = Q/V o , and
R R R R R R

capacitance C with the dielectric present is C = Q/V. The


50
charge Q is the same in both cases, and V is less than V o , so it R R

is concluded that the capacitance C with the dielectric present


30

is greater than C o . R R When the space between plates is


completely filled by the dielectric, the ratio of C to C o (equal R R

to the ratio of V o to V) is called the “dielectric constant” of the


R R

material K:
C
K= (5.5)
Co
When the charge is constant the potential is reduced by a
factor K,
-78-

Vo
K= (5.6)
V
The dielectric constant K is a pure number. Because C
is always greater than C o , K is always greater than unity. A
R R

few representative values of K are given in Table (5.1). For


vacuum, K=1 by definition. For air at ordinary temperatures

0
and pressures, K is about 1.0006; this is so nearly equal to 1

10
that for most purpose an air capacitor is equivalent to one in
vacuum.

40
Table (5.1): Values of dielectric Constant K at 20oC
51 P P

Material K Material K
Vacuum 1.0 Polyvinyl chloride 3.18
50
Air (1 atm) 1.0006 Plexiglas 3.40
Air (100 atm) 1.0548 Glass 5-10
30

Teflon 2.1 Neoprene 6.70


Polyethylene 2.25 Germanium 16
Benzene 2.28 Glycerin 42.5
Mica 3-6 Water 80.4
Mylar 3.1 Strontium titanate 310

When a dielectric material is inserted between the plates


of a capacitor while the charge is kept constant, the potential
-79-

difference between the plates decreases by a factor K.


Therefore, the electric field between the plates must decrease
by the same factor. If E o is the vacuum value and E the value
R R

with the dielectric, then:


Eo
E= (5.7)
K

0
The fact that E is smaller when the dielectric is present

10
means that the surface charge density is smaller. The surface
charge on the conducting plates does not change, but an

40
induced charge of the opposite sign appears on each surface of
the dielectric (Fig. 5.3). These induced surface charge are a
51
result of redistribution of charge within the dielectric material,
a phenomenon called polarization.
50
30

Figure (5.3): Electric field


lines with: a) vacuum
between the plates, and b)
dielectric between the plates.
-80-

It will be assumed in this discussion that the induced


surface charge is directly proportional to the electric field in
the material (like a spring that obeys Hook’s law). In that case,
K is a constant for any particular material. When very strong
fields or anisotropic material are present, the relationship
become more complex, but such cases will not be considered

0
here.

10
A relation between this induced surface charge and the
charge on the plates can be derived. Let us denote the

40
magnitude of the charge per unit area induced on the surfaces
of the dielectric (the induced surface charge density) by σ i ..
51 R R

The magnitude of the surface charge density on the capacitor


plates is σ as usual. Then the net surface charge on each side
50
of the capacitor has magnitude (σ-σ i ) as shown in Fig. 5.3b.
R R

The field between the plates is related to the net surface charge
30

density σ net by:


R R

E=σ net /ε o .
R R R R

Without and with the dielectric respectively, we have:


σ
Eo= , (5.8)
εo
R R

and E= σ − σ i .
εo
-81-

Then,
σ - σi = E εo
R R R

Eo ε o
=
K
Using these expressions in Eqn.(5.8) and rearranging the
result, we find :

0
Eo ε o
σi = σ -

10
R R

K
σ εo
=σ-
εo K

40
σ 1
=σ- = σ (1 − ) (5.9)
k K
51
This equation shows that when K is very large σ i is R R

nearly as large as σ. In this case σ i nearly cancels σ and the


50
R R

field and potential difference are much smaller than their


30

values in vacuum.
The product Kε o is called the permittivity of the
R R

dielectric denoted by ε:
ε = Kε o R R (5.10)

In terms of ε, the electric field within the dielectric can


be expressed as:
E = σ/ε (5.11)
-82-

The capacitance when the dielectric is present is given by:


C = KC o R

A
= Kε o
d
(5.12)
A

d
In empty space where K=1, ε=ε o . For this reason, ε o is

0
R R R R

sometimes called the permittivity of free space, or the

10
permittivity of vacuum. Because K is a pure number, ε and ε o R R R R

have the same units; C2/N m2 or F/m.

40
P P P P

We can repeat the derivation of Eqn. (5.4) for the energy


51
density u in an electric field for the case in which a dielectric is
present. The result is:
50
u = ½ K ε o E2
R R P

= ½ ε E2 (5.13)
30

P P P P
Chapter VI: Moving Charge

6.1. Current:
A current is any motion of charge from one region to

0
another. In the electrostatic situations, which was discussed in

10
the past four chapters, there are no charges in motion, and the
electric field everywhere within a conductor is zero. Now we

40
let the charges move and current in conducting materials will
be discussed. 51
To maintain a steady flow of charge in a conductor, we
have to maintain a steady force on the mobile charges in the
50
conductor. Assume that there is an electric field E within the
conductor, so a particle with charge q experiences a force F =
30

qE.
When a charged particle such as an electron moves in an
electric field in vacuum, it accelerates continuously. However,
the motion of an electron inside a conducting material such as
a metal is very different because of frequent collisions with the
atoms of the material. In a metal the free electrons have a lot
of random motion, somewhat like the molecules in a gas but
with much greater speeds, of the order of 106 m/s.
-84-

When an electric field is applied to the metal, the forces


that it exerts on the electrons lead to a small net motion or drift
in the direction of the force, in addition to this random motion.
The electric field does work on the moving charges, and the
resulting kinetic energy is transferred to the material by means
of inelastic collisions with the ion cores, which vibrate about

0
their equilibrium positions in a crystal lattice. This energy

10
transfer increases their average vibration energy and therefore
the temperature of the material. Thus the motion of the

40
charged particles consists of random motion with very large
average speed and a much slower “drift speed” of the order of
51
10-4 m/s in the direction of the electric-field force.
P P

Current is defined to be “the amount of charge


50
transferred per unit time”. Thus, if a net charge dQ flows
through an area in a time dt the current through the area
30

denoted by i, is:
i = dQ/dt (6.1)
In metals, the moving charges are always negative
electrons, but in an ionized gas (plasma) or an ionic solution,
both electrons and positively charged ions are moving. In a
semiconductor material such as germanium or silicon,
conduction is partly by electron and partly by motion of
-85-

vacancies, also known as holes; these are sits of missing


electrons and act like positive charges.
Figure 6.1 compares the motion of positive and negative
charges. In Fig. 6.1a the moving charges are positive and move
in the direction of the electric field. In Fig. (6.1b) they are
negative and move oppositely to the electric field. In both

0
cases the result is a net transfer of positive charge from left to

10
right.

40
51
50
30

Figure (6.1): Motion of a) positive and b) negative charges in


electric field.
Current is a scalar quantity. The SI unit of current is the
ampere; one ampere is defined as one coulomb per second
(1 A = 1 C/s). This unit is named in honor of the French
scientist Andre Ampere (1775-1836). The current in an
ordinary flashlight (D-cell size) is about 0.5 to 1A; the current
in the starter motor of a car engine is around 200 A. Current in
-86-

radio and television circuits are usually expressed in


milliamperes (1mA=10-3 A) or microamperes (1 µA = 10-6 A)
P P P P

and currents in computer circuits are expressed in


nanoamperes (1 nA=10-9 A) or picoamperes (1 pA = 10-12 A).
P P P P

The current through an area can be expressed in terms of


the drift velocity, v d , of the moving charges. Let us consider a

0
R R

conductor with cross-section area A and an electric field E

10
directed from right to left as shown in Fig. (6.2). The case in
which the free charges in the conductor are negative

40
(electrons) will be discussed; then the electric force is in the
opposite direction for the field. It will be also assumed that the
51
material of the conductor is homogeneous (the same properties
at all points) and isotropic (all directions are equivalent).
50
30

Figure (6.2): Electric field in a conductor.

Suppose there are n charged particles per unit volume.


We call n the concentration of particles; its SI unit is m-3. P P

Assume that all the particles move with a drift velocity with
-87-

magnitude v d . In a time dt, each particle moves a distance


R R

v d dt.
R R

The particles that flow out of the right end of the shaded
cylinder with length v d dt during dt are the particles that were
R R

within this cylinder at the beginning of dt. The volume of the


cylinder is A v d dt, and the number of particles within it is n A

0
R R

v d dt. If each particle has a charge q, the charge dQ flowing

10
R R

out of the end of the cylinder in time dt is:


dQ = q (nA v d dt)
R R

40
and the current is :
dQ
I= = nqv d A. (6.2)
dt
51
50
The current per unit cross-section area is called the
current density, J. From Eqn. (6.2) :
30

I
J= = nqv d (6.3)
A
If the moving charges are positive rather than negative,
the electric force will be in the same direction of E, and the
drift velocity is right to left, as in Fig. 5-1b. But the current is
still left to right; negative charge moving right to left and
positive charge moving left to right would both increase the
positive charge at the right of the section. In either case,
-88-

particles flowing out at an end of the cylindrical section are


continuously replaced by particles flowing in at the opposite
end.
In general, a conductor may contain several different
kinds of charged particles having charges q 1 , q 2 , ..... , R R R R

concentrations n 1 , n 2 , ....., and drift velocities with magnitudes

0
R R R R

v d1 , v d2 , .... , An example is conduction in an ionic solution.

10
R R R R

The total current is then :

40
I = A(n 1 q 1 v d1 + n 2 q 2 v d2 + ... )
R R R R R R R R R R R R (6.4)

When more than one kind of charge is moving, the total


51
current density J = I/A is given by :
50
J = n 1 q 1 v d1 + n 2 q 2 v d2 + ...
R R R R R R R R R R R R

(6.5)
30

A vector current density J can also be defined that


includes the directions of the drift velocities;
J = n 1 q 1 v d1 + n 2 q 2 v d2 + ...
R R R R R R R R R R R R

(6.6)

The direction of the drift velocity v d is the same as that R R

of the electric field E for a positive charge and opposite to E


for a negative charge. In both cases the product qv d is in the R R
-89-

direction of E. The vector current density J always has the


same direction as the field E, even in a metallic conductor, in
which the moving charges are (negative) electrons with v d in R R

the direction opposite to E.


When there is a steady current in a closed loop, the total
charge in every segment of the conductor is constant. From

0
the principle of conservation of charge the rate of flow of

10
charge out at one end of a segment, at any instant equals the
rate of flow of charge in at the other end of the segment, and

40
the current is the same at all cross sections. Current is not
something that squirts out of the positive terminal of a battery
51
and is consumed or used up by the time it reaches the negative
terminal.
50

6.2. Resistance:
U U
30

The current density J in a conductor depends on the


electric field E and on the properties of the material. In
general, this dependence can be quit complex. For some
material, especially metals, J is nearly directly proportional to
E. In that case, the ratio of the magnitudes E and J is constant.
This ratio of the magnitudes of electric field and current
density gives the resistivity ρ of a material that is defined as:
ρ=E/J (6.7)
-90-

It will be assumed in the following discussion that this


simple proportion is valid, even though there are many
situations in which it is not.
The greater the resistivity, the greater the field needed to
cause a given current density or the smaller the current density

0
caused by a given field. From Eqn. (6.7), the units of ρ are

10
(V/m)/(A/m2) = V.m/A. 1 V/A is called one ohm (1 Ω), so
P P

alternative units for ρ are Ω.m (ohm-meters). A “perfect”

40
conductor would have zero resistivity and a “perfect” insulator
would have an infinite resistivity. Metals and alloys have the
51
smallest resistivities and are the best conductors. The
resistivities of insulators are greater than those of the metals by
50
an enormous factor, of the order of 1022.
P P

The proportionality of J to E for a metallic conductor at


30

constant temperature was discovered by G.S. Ohm (1787-


1854) and is called “Ohm’s law”. A material that obeys Ohm’s
law reasonably well is called an ohmic conductor, or a linear
conductor. Many materials show substantial departures from
Ohm’s-law behavior; they are nonohmic, or nonlinear. Ohm’s
law, like the ideal gas equation and Hooke’s law, is an
idealized model that describes the behavior of some materials
-91-

quite well but is not a general description of all matter. The


reciprocal of resistivity is conductivity. Its units are (Ωm)-1. P P

Comparing electrical resistivities with thermal


conductivitys we note that good electrical conductors, such as
metals, are usually also good conductors of heat. Poor
electrical conductors, such as ceramic and plastic materials,

0
are also poor thermal conductors. In a metal the free electrons

10
that carry charge in electrical conduction also provide the
principal mechanism for heat conduction. So we should expect

40
a correlation between resistivities of electrical conductors and
insulators, it is easy to confine electric currents to well-defined
51
paths or circuits. The variation in thermal conductivity is
much less, only a factor 103 or so, and it is usually impossible
P P
50
to confine heat currents to that extent.
Semiconductors have resistivities that are intermediate
30

between those of metals and those of insulators. These


materials are important not primarily because of their
resistivities, but because of the way their resistivities are
affected by temperature and by small amount of impurities.

For a conductor that obeys Ohm’s law, with resistivity


ρ, the current density J at a point where the electric field is E is
given by Eqn. (6.7), which we can write as:
-92-

E=ρJ (6.8)
When Ohm’s law is obeyed, ρ is constant and E is
directly proportional to J. Often we are more interested in the
total current i in a conductor and the potential difference V
between its ends. For example, suppose our conductor is a
wire with uniform cross-section area A and length L, as shown

0
in Fig. (6.3).

10
40
51
50
30

Figure (6.3): Homogeneous conductor with uniform cross-section.

If the magnitudes of the current density J and the


electric field E are uniform throughout the conductor, the total
current I is given by I = JA, and the potential difference V
between the ends is V = E L. When these equations are solved
for J and E, respectively, and substitute the results in Eqn.
(6.8), we obtain:
-93-

V ρI
=
L A
or
ρL
V= I (6.9)
A
This shows that when ρ is constant, the total current I is

0
proportional to potential difference V.

10
The ratio of V to I for a particular conductor is called its
resistance R:

40
V
R= (6.10)
I 51
Comparing this definition of R to Eqn. (6.9), it is seen
that the resistance R of a particular conductor is related to the
resistivity ρ of its material by:
50

ρL
R= (6.11)
30

A
If ρ is constant, then so is R. Equation (6.11) shows that
the resistance of a wire or other conductor of uniform cross
section is directly proportional to its length and inversely
proportional to its cross section area. It is also proportional to
the resistivity of the material of which the conductor is made.
Rearrange Eqn.(6.10), then
V = IR. (6.12)
-94-

This equation is often called “Ohm’s law”, but it is


important to understand that the real content of Ohm’s law is
the direct proportionality (for some materials) of V to I or of J
to E. Equation (6.10) or (6.12) defines resistance R for any
conductor, whether or not it obeys Ohm’s law, but only when
R is constant can we correctly call this relationship Ohm’s law.

0
The SI unit of resistance is the ohm, equal to one volt

10
per ampere (1Ω=1 V/A). The kilohm (1kΩ = 103 Ω) and the
P P

megohm (1MΩ=106 Ω) are also in common use. A 100-W,

40
P P

120-V light bulb has a resistance (at operating temperature) of


140Ω. A 100-m length of 12-gauge copper wire (the size
51
usually used in household wiring) has a resistance at room
temperature of about 0.5Ω. Resistors in the range 0.01Ω to 107
50
P P

Ω can be bought off the shelf.


A circuit device made to have a specific value of
30

resistance is called a resistor. Individual resistors in electronic


circuity are often cylindrical in shape, a few millimeters in
diameter and length, with wires coming out the ends.The
resistance may be marked with a standard code three or four-
color bands near one end (Fig.6.4). The first two bands
(starting with the band nearest an end) are digit, and the third
is a power-of-10 multiplier, as shown in Fig.(6.4). For
-95-

example, yellow-violet-orange means 47 x 103Ω or 47 KΩ.


P P

The fourth band, if present, indicates the precision of the


value; no band means ± 20%, a silver band ± 10% and a gold
band ±5%. An important characteristic of a resistor is the
maximum power that it can dissipate without damage.

0
10
40
51
Figure (6.4): Resistor.
50
For a resistor that obeys Ohm’s law a graph of current
as a function of potential difference (voltage) is a straight line
30

(Fig.6.5a). The slope of the line is I/R. As mentioned before,


not all devices obey Ohm’s law. The relation of voltage to
current may not be a direct proportion, and it may be different
for the two directions of current.
-96-

0
10
Figure (6.5): I-V characteristics.

40
In devices that do not obey Ohm’s law, the relationship
of voltage to current may not be a direct proportion, and it may
51
be different for the two directions of current. Figure (6.5b)
shows the behavior of a semiconductor diode, a device used to
50
convert alternating current to direct current and to perform a
wide variety of logic functions in computer circuitry. For
30

positive potentials V of the anode (one of two terminals of the


diode) with respect to the cathode (the other terminal), I
increases exponentially with increasing V; for negative
potentials the current is extremely small. Thus a positive V
causes a current to flow in the positive direction, but a
potential difference of the other sign causes little or no current.
Hence a diode acts like a one-way valve in a circuit.
Chapter VII: Electromotive Force

7.1. Circuits:
For a conductor to have steady current it must be part of
a path that forms closed loop, or complete circuit. But the path

0
cannot consist entirely of resistance. In a resistor, charge

10
always moves in the direction of decreasing potential energy.
There must be some part of the circuit where the potential

40
energy increases.

51
The problem is analogous to an ornamental water
fountain that recycles its water. The water pours out of
50
openings at the top, cascades down over the terraces and
spouts, and collects in a basin in the bottom. A pump then lifts
30

it back to the top for another trip. Without the pump the water
would just fall to the bottom and stay there.

In an electric circuit there must be a device somewhere


in the loop where a charge travels “uphill,” from lower to
higher potential, despite the fact that the electrostatic force is
trying to push it from higher to lower potential. The influence
that makes charge move from lower to higher potential is
-98-

called electromotive force (emf). Every complete circuit with a


steady current must include some device that provides emf.
This is a poor term because emf is not a force but an energy-
per-unit-charge quantity, like potential.

The SI unit of emf is the same as the unit for potential,

0
the volt (1V= 1 J/C). A battery with an emf of 1.5 V does 1.5

10
J of work on every coulomb of charge that passes through it.
We’ll use the symbol ε for emf.

40
Batteries, electric generators, solar cells, thermocouples,
51
and fuel cells are all examples of sources of emf. All such
devices convert energy of some form (mechanical, chemical,
50
thermal, and so on) into electrical energy and transfer it into
the circuit where the device is connected. An ideal source of
30

emf maintains a constant potential difference between its


terminals, independent of the current through it. We define
emf quantitatively as the magnitude of this potential
difference. As we will see, such an ideal source is a mythical
beast, like the frictionless plane. We will discuss later how real
life sources of emf differ in their behaviour from this idealized
model.
-99-

Figure (7.1a) is a schematic diagram of a source of emf


that maintains a potential difference between terminals ‘a’ and
‘b’ called the terminals of the device. Terminal ‘a’, marked +,
is maintained at higher potential than terminal ‘b’, marked -.

0
10
40
51
50
(a) (b)
Figure (7.1): Schematic diagram of a source.
30

Associated with this potential difference is an electric


field E in the region around the terminals, both inside and
outside the source. The electric field inside the device is
directed from ‘a’ to ‘b’, as shown. A charge q within the
source experiences an electric force F e = qE. The source has to
R R

provide some additional influence, which we represent as a


force F n that pushes charge from ‘b’ to ‘a’ inside the device
R R
-100-

(opposite to the direction of F e ) and maintains the potential


R R

difference. The origin of this additional influence depends on


what kind of source we are talking about. In a generator, the
additional influence results from magnetic-field forces on
moving charges. In a battery or fuel cell, it is associated with
diffusion processes and varying electrolyte concentration

0
resulting from chemical reactions. In an electrostatic machine

10
such as a Van de Graaff generator an actual mechanical force
is applied by a moving belt or wheel.

40
The potential V ab of point a with respect to point b is
R R
51
defined as the work per unit charge done by the electrostatic
force F e =qE on a charge q that moves from a to b. The emf ε
R R
50
of the source is the energy per unit charge supplied by the
source during the “uphill” displacement from b to a. For the
30

ideal source of emf, the potential difference V ab , in case of no


R R

complete circuit, is equal to the electromotive force ε :


V ab = ε
R R (7.1)

Now let us make a complete circuit by connecting a


wire with resistance R to the terminal of a source (Fig. 7.1b).
The charged terminals a and b of the source set up an electric
field in the wire, and this causes a current in the wire from a
-101-

toward b. From V = IR the current I, in case of ideal source


of emf , is determined by:
ε = V ab = IR
R R (7.2)
That is, when a charge q flows around the circuit the
potential rise δ as it passes through the source is numerically
equal to potential drop V ab =IR as it passes through the resistor.

0
R R

Once ε and R are known, this relation determines the current

10
in the circuit. The current is the same at every point in the
circuit. This follows from conservation of charge and from the

40
fact that charge can not accumulate in the circuit devices that
we have described (Otherwise the potential differences would
change with time).
51
50
7.2. Internal Resistance:
U U

Real sources do not behave exactly the way we have


30

described because charge moving through the material of any


real sources encounters resistance. We call this the internal
resistance of the source, denoted by r. If this resistance
behaves according to Ohm’s law, r is constant. The current
through r has an associated drop in potential equal to Ir. The
terminal potential difference V ab under complete-circuit
R R

conditions and source with internal resistance is then:


-102-

V ab = ε - Ir
R R (7.3)

The potential v ab , called the terminal voltage, is less than the


R R

emf because of the term Ir representing the potential drop


across the internal resistance r.
The current in the external circuit is still determine by

0
V = IR.

10
Combining this with Eqn. (7.3), one get:
ε - Ir = IR,

40
or
ε 51
I= (7.4)
R+r
That is, the current equals the source emf divided by the
50
total circuit resistance (R+r). Thus, the behaviour of a source
can be described in terms of two properties: an emf ε, which
30

supplies a constant potential difference independent of current,


in series with an internal resistance r.
To summarize, a circuit is a closed conducting path
containing resistors, sources of emf, and possibly other circuit
elements. Equation (5-17) shows that the algebraic sum of the
potential difference and emf around the path is zero. Also, the
current in a simple loop is the same at every point. Charge is
conserved; if the current were different at different points,
-103-

there would be a continuing accumulation of charge at some


points, and the current couldn’t be constant.

Idealized meters do not disturb the circuit in which they


are connected. An idealized voltmeter has infinitely large
resistance and measures potential difference without having

0
any current diverted through it; an idealized ammeter has zero

10
resistance and measures the current through without having
any potential difference between its terminals.

40
Finally, we remark that Eqn.(7.3) is not always an
51
adequate representation of the behaviour of a source. The emf
may not be constant, what we have described as an internal
50
resistance may actually be a more complex voltage-current
relation that does not obey ohm’s law. Nevertheless, the
30

concept of internal resistance frequently provides an adequate


description of batteries, generators, and other energy
converters. The difference between a fresh flashlight battery
and an old one is not in the emf, which decreases only slightly
with use, but mostly in the internal resistance, which may
increase from a few ohms when fresh to as much as 1000 Ω or
more after long use.
-104-

Similarly, a car battery can deliver less current to the


starter motor a cold morning than when the battery is warm,
not because the emf is appreciably less but because the internal
resistance is temperature-dependent, increasing with
decreasing temperature.

0
7.3. Energy and Power in Electric Circuits:

10
U U

Let’s now look at some energy and power relations in


electric circuits. Figure (7.2) represents a circuit element with

40
potential difference V a -V b = V ab between its terminals and
R R R R R R

current I passing through it in the direction from ‘a’ toward


51
‘b’. This element might be a resistor, a battery, or something
else.
50
30

Figure (7.2): Schematic diagram of circuit element.

As charge passes through the circuit element, the


electric field does work on the charge. In a source of emf,
additional work is done by the force F n that we mentioned in
R R

section (7.1). The total work done on a charge q passing


-105-

through the circuit element is equal to the product of q and the


potential difference V ab (work per unit charge). When V ab is
R R R R

positive, the electric force (and possibly a nonelectric force)


does a positive amount of work qV ab on the charge as it “falls” R R

from potential V a to lower potential V b . If the current is I,


R R R R

then in a time interval dt an amount of charge dQ=I dt passes

0
through. The work dW done on this amount of charge is:

10
dW = V a dQ R R

= V b I dt.
R R

40
This work represents electrical energy transferred into
this circuit element. The time rate of energy transfer is power,
51
denoted by P. Dividing the above equation by dt, we obtain the
rate at which the rest of the circuit delivers electrical energy to
50
this circuit element:
dW
P=
30

dt (7.4)
= Vab I

It may happen that the potential at b is higher than at a;


then V ab is negative, and there is a net transfer of energy out of
R R

circuit element. The element is then acting as a source,


delivering electrical energy into the circuit in which it is
connected. This is the usual situation for a battery, which
-106-

converts chemical energy into electrical energy and delivers it


to the external circuit. Similarly, a generator converts
mechanical energy to electrical energy and delivers it to
external circuit.

The unit of V ab is one volt, or one joule per coulomb,

0
R R

and the unit of I is one ampere, or one coulomb per second.

10
We can confirm that the SI unit of power is one watt:
(I J/C)(I C/s) = I J/s = 1 W.

40
51
50
30
Chapter VIII: Direct Current

8.1. Kirchhoff’s Rules:


Many practical networks cannot be reduced to simple

0
series-parallel combinations. Figure (8.1a) shows a dc power

10
supply with emf ε 1 charging a battery with emf ε 2 and feeding
current to a light bulb with resistance R. Figure (8.1b) is a

40
“bridge” circuit, used in many different types of measurement
and control systems. 51
50
30

Figure (8.1): Different networks


We do not need any new principles to compute the
currents in these networks, but there are some techniques that
help us to handle such problems systematically. We will
describe one of these, first developed by Gustav Robert
Kirchhoff (1824-1887).
-108-

First, here are two terms that we will use often. A


junction in a circuit is a point where three or more conductors
meet. Junctions are also called nodes, or branch points. A
loop is any closed conducting path. The circuit in Fig. (8.1a)
has two junctions; ‘a’ and ‘b’. In Fig.(8.1b), points a, b, c, and
d are junctions, but e and f are not. Some possible loops in

0
Fig.(8.1b) are the closed paths acdba, acdefa, abdefa, and

10
abcdefa.
Kirchhoff’s rules consist of the following two

40
statements:
A) Junction rule: The algebraic sum of the currents into any
51
junction is zero; that is:
∑I= 0 (any junction) (8.1)
50

B) Loop rule : The algebraic sum of the potential differences


30

in any loop, including those associated with emf’s and those of


resistive elements, must equal zero; that is :
∑V = 0 (any closed loop) (8.2)

The junction rule is based on conservation of electric


charge. No charge can accumulate at a junction, so the total
charge entering the junction per unit time must equal the total
-109-

current leaving per unit time. Charge per unit time is current,
so if we consider the currents entering as positive and those
leaving as negative, the algebraic sum of currents into a
junction must be zero. It is like a T branch in a water pipe; if
you have one liter per minute coming in one pipe, you can not
have three liters per minute going out the other two pipes.

0
10
The loop rule is based on the fact that the electrostatic
field is a conservative force field. Suppose we go around a

40
loop, measuring potential differences across successive circuit
elements as we go. When we return to the starting point, we
51
must find that the algebraic sum of these differences is zero;
otherwise, we could not say that the potential at this point has
50
a definite value.
In applying the loop rule, we need some sign
30

conventions. We first assume a direction for the current in


each branch and mark it on the diagram. Then staring at any
point in the circuit, we go around a loop, adding emf’s and IR
products as we come to them. When we go “uphill” through a
source, in the direction from - to +, the emf is considered
positive; when we go “downhill,” from + to -, it is negative.
When we go through a resistor in the same direction as the
assumed current, the IR product is negative because the
-110-

current goes “downhill,” the direction of decreasing potential.


When we go “uphill” through a resistor, in the opposite
direction to the assumed current, the IR term is positive
because this represents a rise of potential.

Kirchoff’s two rules are all that we need to solve a wide

0
variety of network problems. Usually some of the emf’s,

10
currents, and resistance are know, and others are unknown.
We must always obtain from Kirchholff’s rules a number of

40
independent equations to the number of unknowns so that we
can solve the equations simultaneously. Often the hardest part
51
of the solution is not in understanding the basic principle but in
keeping track of algebraic signs.
50

8.2. Charging a Capacitore:


U U
30

In the circuit that we have analyzed up to this point we


have assumed that all the emf’s and resistance are constant
(time-independent) so all potentials currents, and powers have
also been independent of time. But in the simple act of
charging a capacitor we find a situation in which the currents,
voltage and powers do change with time. Figure (8.2) shows a
simple circuit for charging a capacitor. We idealize the battery
(or power supply) to a constant emf ε and zero internal
-111-

resistance r, and we neglect the resistance of all the connecting


conductors.

0
10
Figure (8.2): Circuit of charging a capacitor.

40
We begin with the capacitor initially uncharged; then at
51
some initial time t = 0, the switch is closed, completing the
circuit and permitting current around the loop to begin
50
charging the capacitor. The current begins at the same instant
in every part of the circuit, and at each instant the current is the
30

same in every part. To distinguish between quantities that vary


with time and those that are constant, we will use lower case
letters for time-varying voltage, currents, and capital letters for
constants.
Because the capacitor is initially uncharged, the
potential difference across it is initially zeroed. At this time,
from Kirchhoff’s loop rule,
ε - IR = ε - V ab
R R
-112-

=0
where the voltage V ab across the resistor R is equal to the
R R

battery emf ε. The initial current through the resistor, which


we will call I o is given by Ohm’s law:
R R

I o = V ab /R = ε/R.
R R R R

As the capacitor charges, its voltage V bc increases, and

0
R R

the potential difference V ab across the resistor decreases,

10
R R

corresponding to a decrees in current. From Kirchhoff loop


rule,

40
ε - V ab – V bc = 0
R R R R

or
51
ε = V ab + V bc R R R

Thus, the sum of these two voltages is constant and equal to ε.


50
After a long time the capacitor becomes fully charged, the
current decreases to zero, and the potential difference v ab
30

R R

across the resistor becomes zero. Then the entire battery emf ε
appears across the capacitor, V bc = ε. R R

Let q represent the charge on the capacitor and i the


current in the circuit at some time t after the switch has been
closed. The instantaneous potential difference v ab and v bc are : R R R R

v ab =iR,
R R
-113-

q
or v bc =
R R (8.3)
C
From Kirchhoff’s rule,
q
ε - iR- = 0. (8.4)
C
Solving this equation for i, one get:

0
ε q
i= − (8.5)

10
R RC
At time t = 0, when the switch is first closed, the

40
capacitor is uncharged, so, q = 0. Substituting q = 0 into Eqn.
(8.5), we find that the initial current I o is given by I o = ε/R, as
R R R R

51
we have already noted. If the capacitor were not in the circuit,
the last term in Eqn. (8.5) would not be present; then the
current would be constant and equal to ε/R.
50
As the charge q increases, the term q/RC becomes larger
30

and the capacitor charge approaches its final value, which we


will call Q. The current decreases and eventually becomes
zero. When I= 0, Eqn. (8.5) gives:
ε Qf
= ,
R RC
and then,
Qf = C ε
R R (8.6)
Therefore, the charge Q f does not depend on R.
R R
-114-

The current and the capacitor charge are shown as


functions of time in Fig.(8.3). At the instant the switch is
closed (t=0), the current jumps from zero to its initial value
I o =ε/R; after that it gradually approaches zero. The capacitor
R R

charges starts at zero and gradually approaches the final value


Q =Cε .

0
10
40
51
Figure (8.3): Variation of current and capacitor charge
50
with time.
General expressions can be driven for the charge q` and
30

current i as functions of time. In Eqn. (8.5) we first replace i


by dq`/dt`:

dq `
i=
dt `
ε q`
= −
R RC
1
=− (q ` − Cε )
RC
-115-

This can be arranged to:


dq ` dt
=−
q ` − Cε RC
and then integrate both sides. In this case, q and t are used for
the upper limits and the lower limits are q` = 0 and t` = 0:
q
dq ′ t
dt ′

0
∫ q′ − Cε ∫ RC
= −

10
0 0
then,
t` t
`
− Cε]q0 =−[

40
[ln q ]0
RC
When we carry out the integration and rearrange the
51
result, we get:
q − Cε t
ln[ ]= −
50
− Cε RC
or
q − Cε
30

= e − t / RC
− Cε
Rearrange and use Eqn.(8.6), one get,
q = Cε(1-e-t/RC) P P

= Q f (1- e-t/RC)
R R P P (8.7)
The instantaneous current i is just the time derivative of
Eqn.(8.7):
-116-

dq
i=
dt
ε
= e − t / RC (8.8)
R
= I o e − t / RC
The charge and current are both exponential function of

0
time. Figure (8.3a) is a graph of Eqn. (8.8), and Fig. (8.3b) is a

10
graph of Eq. (8.7).

40
After a time equal to RC the current has decreased to I/e
(about 0.368) of its initial value. At this time the capacitor
charge has reached (1-1/e) = 0.632 of its final value Q f = CV.
51 R R

The product RC is, therefore, a measure of how quickly the


capacitor charges; the product RC has units of time. We call
50
RC the time constant, or the relaxation time, of the circuit, and
is denoted by τ.
30

When τ is small, the capacitor charges quickly; when it


is larger, the charging takes more time. A capacitor charges
more quickly through a small resistor; it is easier for the
charge to get through. In Fig. (8.3a) the horizontal axis is an
asymptote for the curve. Strictly speaking, i never becomes
precisely zero, but the longer one waits, the closer it gets.
-117-

After a time equal to 10RC the current has decreased to


0.000045 of its initial value. Similarly, the curve in Fig. (8.3b)
approaches the horizontal broken line labeled Q f as an R R

asymptote. The charge q never attains precisely this value, but


after a time equal to 10RC, the difference between q and Q f is R R

only 45 millionth’s of Q f .

0
R R

10
8.3. Discharging a Capacitor:
U U

Now suppose that after the capacitor in Fig. 6-2 has

40
acquired a charge Q o . We remove the battery from the circuit
R R

and connect points ‘a’ and ‘c’ (Fig. 8.4).


51 We reset our
stopwatch so that the connection is made at time t = 0 at that
time q = Q o . R R The capacitor then discharges through the
50
resistor, and its charge eventually decreases to zero.
30

Figure (8.4): Circuit of discharging a capacitor.


-118-

Again, let i and q represent the time-varying current and


charge at some instant after the connection is made.
Kirchholff’s loop rule now gives Eqn. (8.5) but with ε = 0 that
is:
dq
i=
dt
(8.10)

0
q
=−

10
RC
The current i is now negative (opposite to the direction

40
shown in Fig. (8.4). At time t = 0 when q = Q o the initial R R

current I o is given as:


R R

I o = - Q o /RC.
R R R R
51
To find q as a function of time, Eq. (6-10) is rearranged,
again change the names of the variables to q` and t`, and
50
integrate. This time the limits for q` are from Q o to q. Then: R R

q
1 t
30

dq`
∫ q`
=− ∫ dt `
RC o
Qo

and
q 1
ln =−
Qo RC′

-t/RC
q=Q o eR R P P (8.11)
-119-

The instantaneous current i is the derivative of this:


dq Q
= − o e − t / RC
dt RC
ε
= − e − t / RC
R
Using Eqn.(8.10), then
-t/RC

0
i = I0 e R R P P (8.12)

10
Both the current and the charge decrease exponentially
with time. Comparing these results with Eqns. (8.7) and (8.-8),

40
we note that the expressions for the current are identical,
except for the sign of I o . The capacitor charge approaches zero
R R

51
asymptotically in Eqn. (8.11), but the difference between q and
Q f approaches zero asymptotically in Eqn. (8.7).
R R
50
Energy considerations give us additional insight into the
behaviour of an RC circuit. While the capacitor is charging,
30

the instantaneous rate at which the battery delivers energy to


the circuit is P = εi. The instantaneous rate at which electrical
energy is dissipated in the resistor is i2R, and the rate at which
P P

energy is stored in the capacitor is iv bc = iq/C. Multiplying R R

Eqn. (8.4) by i, one finds:


εi = i2R + iq/C
P P (8.13)
-120-

This means that of the power εI supplied by the battery


part (i2R) is dissipated in the resistor and part (iq/C) is stored
P P

in the capacitor. The total energy supplied by the battery


during charging of the capacitor equals the battery emf ε
multiplied by the total charge Q f or εQ f . From Eqn. (5.3), the
R R R R

total energy stored in the capacitor is εQ f /2. Thus, the energy

0
R R

supplied the battery is exactly half that is stored in the

10
capacitor, and the other half is dissipated in the resistors. It is
a little surprising that this half-and-half division of energy

40
doesn’t depend on C, R, or ε . This result can also be verified
in detail by taking the integral over time of each of the power
51
quantities mentioned above. We leave this calculation for
your amusement.
50
30
Chapter IX: Magnetism

9.1. Introduction:

0
Magnetic phenomena were first observed at least 2500

10
years ago in fragments of magnetized iron that are found near
Ancient City Magnesia (now Manisa). It was discovered that

40
when an iron rod is brought in contact with a natural magnet,
the rod also becomes magnetized.
51 When such a rod is
suspended by a string from its center, it tends to line itself up
in a north-south direction like a compass needle. Magnets
50
have been used for navigation at least since the eleventh
century.
30

Before the relation of magnetic interactions to moving


charges was understood, the interactions of bar magnets and
compass needles were described in terms of magnetic poles.
The end of a bar magnet that points north is called a north
pole, or N-pole, and the other end is a south pole, or S-pole.
Two opposite poles attract each other, and two like poles repel
each other. The concept of magnetic poles is of limited
-121-
usefulness and is somewhat misleading. There is no evidence
that a single isolated magnetic pole exists; poles always appear
in pairs. If a bar magnet is broken in two, each broken end
becomes a pole. The existence of an isolated magnetic pole, or
magnetic monopoly, would have sweeping implications for
theoretical physics. Extensive searches for magnetic

0
monopolies have been carried out, so far without success.

10
A compass needle points north because the earth is a

40
magnet: its geographical North Pole is close to a magnetic
South Pole. The earth’s magnetic axis is not quite parallel to
51
its geographic axis (the axis of rotation), so a compass reading
deviates somewhat from geographic north; this deviation,
50
which varies with location, is called magnetic declination.
Also the magnetic field is not horizontal at most points on the
30

earth’s surface; its inclination upon or down is described by


the angle of dip.

In 1819 the Danish scientist Hens Christian Oersted


discovered that a compass needle was deflected by a current-
carrying wire. Similar investigations were carried out in
France by André Amper. A few years later, Michael Faraday
in England and Joseph Henry in United States discovered that
-122-
moving a magnet near a conducting loop can cause a current in
the loop and that a charging current in one conducting loop can
cause a current in a separate loop. These observations were
the first evidence of the relationship of magnetism to moving
charges. It is well know now that electric and magnetic
interactions are intimately intertwined.

0
10
9.2. Magnetic Field of Moving Charge:

40
Let us start with the magnetic field of a single moving
point charge q. The location of the charge is called “the source
51
point” and the point P where we want to find the field is called
“the field point”. In the study of electric fields in Chapter 2, it
50
was found that the E field of a point charge q, at a field point a
distance r from the charge, is proportional to q and to 1/r 2. The
30

direction of this field (for positive q) is along the line from


source point to field point. The corresponding relationship for
the magnetic field B of a point charge q moving with velocity
v has some similarities to this relationship and some
interesting differences. Experiments show that the magnitude
B is also proportional to 1/r2 but there the similarity ends. The
direction of B is not along the line from the source point (the
moving charge) to the field point. Instead, B is perpendicular
-123-
to the plane containing this line and the particle’s velocity
vector v, as shown in Fig. (9.1). Furthermore, the field
magnitude B is proportional to the sine of the angle  between
these two directions. Finally, B is proportional to the
particle’s speed v. Thus, the magnitude B of the magnetic
field at point P is given by:

0
 o qv sin 

10
B= (9.1)
4 r 2

40
where o/4 is a proportionality constant. The reason for
writing the constant in this particular way will emerge shortly.
51
50
30

Figure (9.1): Field lines of moving charge.

-124-
We did something similar with Coulomb’s law. We can
incorporate both the magnitude and direction of B into a single
vector equation using the vector product. To avoid having to
say “the direction from the source q to the field point P”

over and over, a unit vector r = r/r that points in the
direction from charge q to point P, that is, from the source

0
point to the field point. Then the B field of a moving point

10
charge is:
_ 
− o q  xr

40
=
B 4 r 2 (9.2)
51
Figure (9.1) shows the relation of r to P and also shows
the magnetic field B at several points in the vicinity of the
50
charge. At all points along a line through the charge parallel to
the velocity v, the field is zero because sin  = 0 at all such
30

points. At any distance r from q, B has its greatest magnitude


at points lying in the plane perpendicular to v because at all
such points,  = 90, and sin  = 1. The charge also produces
an electric field in its vicinity; the electric-field vectors are not
shown in the figure. The field lines for the electric field of a
positive point charge radiate outward from the charge. The
magnetic field lines are completely different. The above

-125-
discussion shows that for a point charge moving with velocity
v, the magnetic field lines are circles with centers along the
line of v and lying in planes perpendicular to this line. The
directions of these lines for a positive charge are given by a
right-hand rule: Grasp the velocity vector v with your right
hand so that your right thumb points in the direction of v: your

0
fingers then curl around the line of v in the same sense as the

10
magnetic field lines.

40
Figure (9.1a) shows parts of a few field lines, and
Fig.(9.1b) shows some field lines in a plane through q,
51
perpendicular to v, as seen looking in the direction of v. The
unit of B is one Tesla (T):
50
T = N.s/C.m = N/A.m.
30

Using this with Eqn. (9.1) or (9.2) and solving for o, we find
that the units of the constant o are:
1 N.s2/C2= 1 N/A2= 1 Wb/A.m= 1 T.m/A
In SI units the numerical value of o is exactly 4x10-7. Thus
o = 4x10-7 N.s2/C2= 4x10-7 N/A2
= 4x10-7 Wb/A.m= 4x10-7 T.m/A (9.3)

-126-
This numerical value actually stems from the definition
of the coulomb. The constant 1/4o in Coulomb’s law is
related to the speed of light c:
1/4o = (4x10-7 N.s2/C2)c2

When we study electromagnetic waves, it will be found

0
that their speed of propagation in vacuum, which is equal to

10
the speed of light c, is given by:
1

40
C2 = (9.4)
 o o

If the preceding equation is solved for o, substitute the


51
resulting expression into Eqn.(9.4) and solve for o, we get the
value of o stated above. This discussion is a little premature,
50
but it may give you a hint about one of the nice unifying
threads that runs through electromagnetic theory.
30

The magnetic field, like the electric field, obeys the


superposition principle: The total magnetic field caused by
several moving charges is the vector sum of the fields caused
by the individual charges. In the next section, this fact will be
used to calculate the field caused by a current in a conductor.

-127-
9.3. Magnetic Field of Current Element:

We can use the principles of Section 9.2, including the


superposition principle, to find the magnetic field at any field
point P produced by a current in a conductor. The total field is
the vector sum of the fields due to all of the moving charges in

0
the conductor.

10
40
51
50
30

Figure (9.2): Current-carrying conductor.

The magnetic field caused by a short segment dL of a


current-carrying conductor, which is shown in Fig. (9.2a), has
to be calculated. The volume of the segment is A dL, where A
is the cross-section area of the conductor. If there is n charge

-128-
q per unit volume, then the total moving charge dQ in the
segment is:
dQ = nqA dL.

The moving charges in this segment are equivalent to a


single charge dQ, traveling with a velocity equal to the drift

0
velocity, vd. (Fields due to the random motions of the carriers

10
will, on the average, cancel out at every point.). From Eqn.
(9.1) the magnitude of the resulting field dB at any point is:

40
 o dQv d sin 
dB =
4 r2 51
 nqv d Adl sin
= o
4 r2
50
However, nqvdA equals the current I in the element, so :
 o Idl sin 
dB = (9.5)
4 r 2
30

or in vector form,

−  I dl x r
dB = o (9.6)
4 r 2
where dl is a vector with length dl, in the same direction as the
current in the conductor.

-129-
Equations (9.5) and (9.6) are called the Law of Biot and
Savart. To find the total magnetic field B at any point in space
due to the current in a complete circuit, we have to integrate
one of these expressions:

 o Idlxr
B=  r2

0
(9.7)
4

10
In the following, this vector integration will be carried out for
several examples.

40
As shown in Fig. (9.2a), the field vectors dB and the
magnetic-field liens are exactly like those set up by a positive
51
charge dQ moving in the direction of the drift velocity v d. The
field liens are circles in planes perpendicular to dl and centered
50
on the line of dl. Their directions are given by the same right-
hand rule that is introduced for point charges in Section 9.1.
30

Figure (9.2b) shows field lines in a plane containing the


element dl and perpendicular to it; note the similarity to
Fig.(9.1b) for a moving point charge.

One can not verify Eqn. (9.6) directly because we can


never experiment with an isolated segment of a current-
carrying circuit. What we measure experimentally is the total
B for a complete circuit. But we can still verify these
-130-
equations indirectly by calculating B for various current
configurations and comparing the results with experimental
measurements. We can also verify Eqn.(9.1), the starting point
of the derivation.

If matter is present in the space around a current-

0
carrying conductor, the field at a field point P in its vicinity

10
will have an additional contribution resulting from the
magnetization of the material. However, unless the material is

40
iron or some other ferromagnetic material, the additional field
is so small that it is usually negligible. Additional
51
complication arises if time-varying electric or magnetic fields
are present or if the material is a superconductor.
50

9.4. Magnetic Field of Straight Conductor:


30

An important application of the law of Biot and Savart


is finding the magnetic field produced by a straight conductor
with length 2a carrying a current I, at a point on its
perpendicular bisector, at a distance x from the conductor. The
situation is shown in Fig.(9.3). We first use the law of Biot and
Savart, Eqn. (9.5), to find the field dB caused by the element
of conductor dl = dy shown in the figure. From the figure,
-131-
r= x 2 + y 2

and
sin  = sin( - )

= x/ x 2 + y 2 .

From the right-hand rule or the vector product d l x r , the

0
10
direction of dB is perpendicular to the plane of the figure, into
the plane. In this case the directions of the dB’s from all
elements of the conductor are the same. Thus, in integrating

40
Eqn. (9.5), we can just add the magnitudes of the dB’s.
51
50
30

Figure (9.3): Magnetic field produced by strait


current-carrying conductor.

-132-
Putting the pieces together, we find that the magnitude
B of the total B is:
 o I +a xdy
B= 
4 - a (x 2 + y 2 ) 3 / 2

0
This can be integrated by trigonometric substitution or by

10
using an integral table.
+a

40
 I y
B= o
4x ( x 2 + y 2 )
−a
51
The final result is:
oI 2a
B= (9.8)
4 x x 2 + a 2
50
30

When the length (2a) of the conductor is very great in


comparison to its distance x from point P, we can consider it to
be infinitely long.

When ‘a’ is much larger than x, x2 + a a is


approximately equal to a, and so in the limit as a → , Eqn.
(9.8) becomes:

-133-
oI
B=
2x
The physical situation has axial symmetry about the y-
axis, so B must have the same magnitude at all points on a
circle centered on the conductor and lying in a plane
perpendicular to the conductor and its direction is everywhere

0
tangent to such a circle. Thus, at points on a circle of a radius r

10
around the conductor, the magnitude B is given by;
oI
B= (9.9)

40
2 r

51
Part of the magnetic field around a long, straight
conductor is shown in Fig.(9.4). The configuration of the
50
magnetic field lines in this situation is completely different
from that of the electric field lines in the analogous electrical
30

situation. Electric field lines radiate outward from a positive


line charge distribution (inward for negative charges). By
contrast, these magnetic field lines encircle the current that
acts as their source. Electric field lines begin and end at
charge, but magnetic field lines are always continuous loops
and never have end points, irrespective of the shape of the
conductor that sets up the field.

-134-
0
10
40
Figure (9.4):Magnetic field around straight conductor.
51
This property of magnetic fields can be discussed along
with the fact that the total magnetic flux out of a closed surface
50
is always zero. Gauss’s law for magnetic fields,
30

 B. dA = 0 (9.10)

expresses the fact that there are no isolated charges or


magnetic monopoles. The number of magnetic field lines
emerging from any closed surface must equal the number
entering it.

-135-
9.5. Magnetic Field of Circular Loop:

If you look inside a doorbell, a relay, a transformer, or


an electric motor, you will find coils of wire, often consisting
of many circular loops. A current in such a coil is used to
establish a magnetic field. So, it is worthwhile to derive an

0
expression for the magnetic field produced by a single circular

10
conducting loop carrying a current or by N closely spaced
circular loops forming a coil.

40
Figure (9.5) shows a circular conductor with radius a,
51
carrying a current I. The current is led into and out of the loop
through two long, straight wires side the currents in these
50
straight wires are in opposite directions, and their magnetic
fields very nearly cancel each other.
30

Figure (9.5): Magnetic field of circular loop.

-136-
The law of Biot and Savart, Eqn.(8-5) or (8-6), can be
use to find the magnetic field at a point P on the axis of the
loop, at a distance x from the center. As the figure shows, dl
and r are perpendicular, and the direction of the field dB
caused by this particular element dl lies in the xy-plane. Also,
r2 = x2 + a2. The magnitude dB of the field due to element dl

0
is:

10
oI dl
dB = (9.12)

40
4 ( x 2 + a 2 )
51
The component of the vector dB are:
dBx = dB cos 
50
oI dl a
= (9.13)
4  ( x + a ) ( x + a 2 ) 1/2
2 2 2
30

and
dBy = dB sin 
oI dl a
= (9.14)
4  ( x + a ) ( x + a 2 ) 1/2
2 2 2

Because the situation has rotational symmetry about the


x-axis, there cannot be a component of B perpendicular to that
axis. For every element dl there is a corresponding element on
-137-
the opposite side of the loop with opposite direction. These
two elements give equal contribution to the x-component of
the field, given by Eq. (8-13), but opposite components
perpendicular to the x-axis. Thus all the perpendicular
components cancel, and only the x-components survive. To
obtain the total x-component, we integrate Eqn.(8-13),

0
including all the dl’s around the loop. Everything in this

10
expression except dl is constant and can be taken outside the
integral, so we have:

40
B x =  dB x
oI adl
= 
51
4 (x + y 2 ) 3 / 2
2

oI a
=  dl
50
4 ( x 2 + y 2 ) 3 / 2
30

The integral of dl is just the circumference of the circle,

 dl = 2a , and we finally get:

oI a2
Bx= (9.15)
4  (x 2 + y 2 ) 3/ 2

-138-
Now, suppose that instead of the single loop in Fig.
(9.6) we have a coil consisting of N closely spaced loops, all
with the same radius. Each loop contributes equally to the
field, and the total field is N times the field of a single loop :
 o Ia 2 N
Bx = (9.16)
2(x 2 + a 2 ) 3/ 2

0
10
At the center of the loop or loops, x=0, and Eqn.(9.16)
reduces to:

40
oI N
Bx = (9.17)
2a 51
As we go out along the axis, the field decreases in
50
magnitude. Figure (9.6) shows a graph of B x as a function of
x.
30

Figure (9.6): Bx versus x.

-139-
Some of the magnetic field lines surrounding a circular
loop, in planes through the axis, are shown in Fig. (9.7). The
field lines encircle the conductor, and their directions are given
by the right-hand rule, as for a long, straight conductor. Grab
the conductor with your right hand, with your thumb in the
direction of the current; your fingers curl around in the same

0
direction as the field lines. The field lines for the circular loop

10
are not circles, but they are closed curves that enclose the
conductor.

40
51
50
30

Figure (9.7): Magnetic field lines surrounding circular loop.

-140-
Chapter X: Magnetic Field and
Magnetic Force

10.1. Magnetic Field:

0
10
To introduce the concept of magnetic field, let us review
our formulation of electrical interaction in Chapter 2, where

40
we introduced the concept of electric field. We represented
electrical interactions in two steps:
51
1- A distribution of electric charge at rest creates an electric
field E in the surrounding space.
50
2- This electric field exerts a force F = qE on another charge
q that is present in the field.
30

Magnetic interactions can be described in the same way:


A) A moving charge or a current creates a magnetic field
in the surrounding space (in addition to its electric
field).
B) The magnetic field exerts a force F on any other
moving charge or current that is present in the field.

-141-
Like the electric field, the magnetic field is a vector
field, that is, a vector quantity associated with each point in
space. In this chapter we shall concentrate on the second
aspect of the interaction: Given the presence of a magnetic
field, what force does it exert on a moving charge or a current?
In Chapter 9 the problem of how magnetic fields are created

0
by moving charge and current are discussed.

10
What are the characteristics of the magnetic force on a

40
moving charge? First, its magnitude is proportional to the
charge. If a 1-C charge and a 2-C charge move through a
51
given magnetic field with the same velocity, the force on the
2-C charge is twice as great as that on the 1-C charge. The
50
force is also proportional to the magnitude, or “strength” of the
field; if the magnitude of the field is doubled without changing
30

the charge or its velocity, the force doubles. The magnetic


force is also proportional to the particle’s speed. This is quite
different from the electric-field force, which is the same
whether the charge is moving or not.

A charged particle at rest experiences no magnetic force.


Furthermore the “magnetic force” F does not have the same
direction as the magnetic field B but instead is always
-142-
“perpendicular to both B and v”. The magnitude F of the force
is found to be proportional to the component of v
perpendicular to the field; when that component is zero (that
is, when v and B are parallel or antiparallel), the force is zero.
Figure (10.1) shows these relationships. The direction of F is
always perpendicular to the plane containing v and B. Its

0
magnitude is given by:

10
F = q v⊥ B = q vBsin  (10.1)

40
where q is the magnitude of the charge and  is the angle
51
measured from the direction of v to the direction of B, as
shown in the Fig.(10.1).
50
30

Figure (10.1): Directions of moving charge, magnetic


field and magnetic force.
This description does not specify the direction of F
completely; there are always two directions, opposite to each
-143-
other and both perpendicular to the plane of and B. To
complete the description, we use the same right-hand rule that
we used to define the vector product (It would be a good idea
to review that section before you go on). Draw the vectors v
and B with their tails together. Imagine turning v until it
points in the direction of B (turning through the smaller of the

0
two possible angles). Think of warping the fingers of your

10
right-hand around the line perpendicular to the plane of v and
B, as in Fig. (10.1). So that, they curl around with this sense of

40
rotation from v to B; your thumb then points in the direction of
the force F on a positive charge. Alternatively, the direction of
51
the force F on a positive charge is the direction in which a
right-hand-thread screw would advance if turned same way.
50

This discussion shows that the force on a charge ‘q’


30

moving with velocity ‘v’ in a magnetic field ‘B’ is given, both


in magnitude and in direction, by:

F = q(v x B) (10.2)
This is the first of several vector products that we will
encounter in our study of magnetic-field relationships.
Eqn.(10-2) is valid for both positive and negative charges.

-144-
When q is negative, the direction of the force F is opposite to
that of (v x B).

If two charges with equal magnitude and opposite sign


move in the same magnetic field B with the same velocity
(Fig. 10.2), the forces have equal magnitude and opposite

0
direction. Figures (10.1) and (10.2) show several examples of

10
the directions of F and B for both positive and negative
charges.

40
51
50
30

Figure (10.2): Moving two charges.

Equation (10.1) can be interpreted in a different but


equivalent way. Recalling that  is the angle between the
directions of vectors v and B, we may interpret B sin  as the
-145-
component of B perpendicular to v that is, B ⊥. With this
notation the force expression becomes:

F = q v B⊥ (10.3)

This form is equivalent to Eqn. (10.1), but is sometimes

0
more convenient to use, especially in problems involving

10
currents rather than individual particles. We will discuss forces
on currents later in this chapter.

40
From Eq. (10.1) the units of B must be the same as the
51
units of F/qv. Therefore, the SI unit of B is equivalent to 1 N.
s/C. m, or, since one coulomb per second (1A = 1C/s),
50
1N/A.m. This unit is called the Tesla (abbreviated T), in
honor of Nikola Tesla (1857-1943), the prominent Serbian-
30

American Scientist and Inventor:


T = N.s/C.m
= N /A.m.
The cgs unit of B, the gauss (1G = 10 -4T), is also in
common use. Instruments for measuring magnetic field are
sometimes called gauss-meters.

-146-
The magnetic field of the earth is of the order of 10 -4T,
or 1G. Magnetic field of the order of 10 T occurs in the
interior of atoms and are important in the analysis of atomic
spectra. The largest values of steady magnetic field that have
been achieved in the laboratory are of the order of 30 T. Some
pulsed-current electromagnets can produce fields of the order

0
of 120 T for short time intervals of the order of a millisecond.

10
The magnetic field at the surface of a neutron star is believed
to be of the order of 108T.

40
To explore an unknown magnetic field, we can measure
51
the magnitude and direction of the force on a moving test
charge. The electron beam in a cathode-ray is a convenient
50
device for making such measurements. The electron gun shots
out a narrow beam of electrons at a known speed. If there is no
30

force to deflect the beam, it strikes the center of the screen. In


principle, an old TV set with the deflection coils disconnected
could be used for the same purpose.
In general, when a magnetic field is present, the electron
beam is deflected. However, if the beam is parallel or
antiparallel to the field, then in Eqn.(10.1),  = 0 or  and F =
0; there is no force and no deflection. If we find that the

-147-
electron beam is undeflected when its direction is parallel to
the y-axis, the B vector must point either up or down.

When we turn the tube 90o, so that its axis is along the
x-axis the beam deflection is corresponding to a force
perpendicular to the plane of B and v. We can perform

0
addition experiments in which the angle between B and v is

10
between zero and 90o to confirm Eqn.(10.1) or (10.3) and the
accompanying discussion. Science, the electron has a negative

40
charge, the force is opposite in direction to the force on a
positive charge. 51
When a charged particle moves through a region of
50
space where both electric and magnetic fields are present, both
fields exert forces on the particle. The total force F is the
30

vector sum of the electric and magnetic forces:


F = q[E + ( v x B)] (10.4)

10.2. Magnetic Field Line and Flux:

Any magnetic field can be represented by lines. The


idea is the same as for the electric field lines that was
introduced in Section 2.2. We draw the lines so that the line
-148-
through any point is tangent to the magnetic field vector B at
that point. Also, we draw the number of lines per unit area
(perpendicular to the area at a given point) to be proportional
to the magnitude of the field at that point.

These lines are called magnetic field lines. They are

0
sometimes called magnetic lines of force, but that is not a good

10
name for them because unlike electric field lines, they do not
point in the direction of the force on a charge. Magnetic field

40
lines do have the direction that a compass needle would point
at echo location; this may help you to visualize them. Just as
51
with electric field lines, we draw only a few representative
lines, otherwise, the lines would fill up all of the space. Also,
50
because the direction of B at each point is unique, field lines
never intersect.
30

The magnetic field of the earth resembles the field of a


bar magnet, with the magnet’s axis tilted with respect to the
earth’s axis of rotation. The earth’s field thought to be caused
by currents in its molten core, changes with time. There is
geologic evidence that it has actually changed direction several
times during the past 100 million years.

-149-
The magnetic flux B through a surface is defined just as
electric flux is defined in connection with Gauss’s law. Any
surface can be divided into elements of area dA (Fig.10.3).

0
10
40
Figure (10.3): Magnetic flux through surface.
51
For each element we determine B⊥, the component of B
50
normal to the surface at the position of that element, as shown.
From the figure, B⊥ = B cos , where  is the angle between
30

the direction of B and a line perpendicular to the surface. (Be


careful not to confuse  with B). In general, this component
varies from point to point on the surface. The magnetic flux
dB through area dA is defined as:
dB = B⊥dA
= B cos  dA
= B. dA. (10.5)

-150-
The total magnetic flux through the surface is the sum of
the contributions from the individual area elements:
 B =  B ⊥ dA
(10.6)
=  B cos dA =  B.dA

When the magnetic field lines are drawn, the number of lines

0
passing through any surface is proportional to the magnetic

10
flux through the surface.

40
Magnetic flux is a scalar quantity. In the special case in
which B is uniform over a plane surface with total area A, B ⊥
51
and  are the same at all points on the surface, and:
50
B = B A cos  (10.7)
If B happens to be perpendicular to the surface, cos  = 1 and
30

Eqn.(10.7) reduces to:


B = BA.

According to Gauss’s law, the total electric flux out of a


closed surface is proportional to the total electric charge
enclosed and to the number of electric field lines coming out
of the surface. Electric field lines begin and end on electric

-151-
charges. For example, if the surface encloses an electric
dipole, the total flux is zero because the total charge is zero. If
there were such a thing as a single magnetic charge (magnetic
monopole), the total magnetic flux out of a closed surface
would be proportional to the total magnetic charge strength
enclosed. It was mentioned that no magnetic monopole has

0
ever been observed, despite intensive searches. We conclude

10
that the total magnetic flux out of a closed surface is always
zero, symbolically, for closed surface,

40
 B.dA = 0 (10.8)
51
This equation is sometimes called Gauss’s law for
magnetism. Electric monopoles (single electric charges) exist,
50
but as far as we know, magnetic monopolies do not. It also
follows from Eqn.(10.8) that magnetic field lines are always
30

continuous. They never have end points, such a point would


indicate the existence of monopole.

The general definition of magnetic flux through an open


surface, Eqn. (10.6), has an ambiguity of sign because of the
two possible choices of direction for the vector area element
dA. For Gauss’s law, which always deals with closed surface,

-152-
dA always points out of the surface. However, some
applications of magnetic flux involve an open surface with a
boundary line. In these cases we choose one of the possible
sides of the surface to be the positive side and use that choice
consistently.

0
The SI unit of magnetic flux is equal to the unit of

10
magnetic field (T) times the unit of area (m 2). This unit is
called the weber (Wb), in honor of Wilhelm Weber (1804-

40
1891). It is given as:
Wb = T.m2 51
Also,
T = N/A.m,
50
so,
Wb = T.m2
30

= N.m/A
If the element of area dA in Eqn. (10.5) is at right angles to the
field lines, then B⊥ = B, calling this area dA⊥, then:
d B
B=
dA ⊥
That is, the magnitude of magnetic field is equal to flux per
unit area across an area at right angles to the magnetic field.

-153-
That is why, magnetic field B is sometimes called magnetic
flux density.

10.3. Motion of Charged Particles in Magnetic Field:

0
When a charged particle moves in a magnetic field, the

10
motion is determined by Newton’s laws, with the magnetic
force given by Eqn. (10.2). Figure (10.4) shows a simple

40
example of a particle with positive charge q at point O,
moving with velocity v in a uniform magnetic field B directed
51
into the plane on the figure. The vectors v and B are
perpendicular, so the magnetic force is given as:
50
F = q (v x B).
It has magnitude F = qvB and its direction is as shown in
30

Fig.(10.4). The force is always perpendicular to v, so it can not


change the magnitude of the velocity, only its direction. To put
it differently, the force can not do work on the particle, so the
force can not change the particle’s kinetic energy. Thus, the
magnitudes of both F and v are constant.

-154-
0
10
Figure (10.4): Charged particles move in magnetic field.

40
At points such as P and S in Fig.(10.4a) the directions of
51
force and velocity have changed as shown, but their
magnitudes are the same. The particle, therefore, moves under
50
the influence of a constant-magnitude force that is always at
right angles to the velocity of the particle. Comparing these
30

conditions with the discussion of circular motion, we see that


the particle’s path is a circle, traced out with constant speed v
(Fig.10.4). The centripetal acceleration is v 2/R, and from
Newton’s second law:
F= qvB

v2
= m. (10.9)
R

-155-
where m is the mass of the particle. The radius R of the
circular path is:
mv
R= (10.10)
qB

In terms of the magnitude of the particle’s momentum P

0
= mv, we can also write this as R = P/ q B. If the charge q is

10
negative, the particle moves clockwise around the orbit in
Fig.(10.4).

40
The angular velocity of the particle is given by:
=v/R.
51
Combining this with Eqn. (10.10), one gets:
50
v qB qB
= =v = (10.11)
R mv m
30

The number of revolutions per unit time is /2. This


frequency is independent of the radius R of the path. It is
called the cyclotron frequency; in a particle accelerator called
a cyclotron, particles moving in nearly circular paths are given
a boost twice each revolution, increasing their energy and their
orbital radii but not their angular velocity. Similarly, a
magnetron (a common source of microwave radiation for

-156-
microwave ovens and radar systems) emits radiation with a
frequency equal to the frequency of circular motion of
electrons in a vacuum chamber between the poles of a magnet.

If the direction of the initial velocity is not


perpendicular to the field, the velocity component parallel to

0
the field is constant (because there is no force parallel to the

10
field) and the particle moves in a helix (Fig. 10.5). The radius
of the helix is given by Eqn.(10.10), where v is now the

40
component of velocity perpendicular to the B field.
51
50

Figure (10.5): Charged particle moves in helix.


30

Motion of a charged particle in a non-uniform magnetic


field is more complex. Figure (10.6) shows a field produced
by two circular coils separated by some distance. Particles near
either coil experience a magnetic force toward the center of the
region; particles with appropriate speeds spiral repeatedly from
one end of the region to the other and back. Because charged
particles can be trapped in such a magnetic field, it is called a
magnetic bottle. This technique is used to confine very hot
-157-
plasma having temperatures of the order of 10 6K. In a similar
way the earth’s non-uniform magnetic field traps charged
particles coming from the sun in doughnut-shaped regions
around the earth. These regions, called the Van Allen radiation
belts, were discovered in 1958 through observation made by
instruments aboard the Explorer I satellite.

0
10
40
51
50
Figure (10.6): Field produces by two circular coils.
30

The magnetic force acting on a charged particle can


never do work because at every instant the force is
perpendicular to the velocity. A magnetic force can change the
direction of motion, but it can never increase or decrease the
magnitude of the velocity. Motion of a charged particle under
the action of a magnetic field alone is always motion with
constant speed.

-158-
10.4. Magnetic Force on Current-Carrying Conductor:

What makes an electric motor work? The forces that


make it turn are forces that a magnetic field exerts on a
conductor carrying a current. The magnetic forces on the
moving charges within the conductor are transmitted to the

0
material of the conductor, and the conductor as whole

10
experiences a force distributed along its length. The moving-
coil galvanometer also uses magnetic forces on conductors.

40
We can compute the force on a current-carrying conductor
starting with the magnetic-field force F = q (v x B) on a single
51
moving charge.
Figure (10.7) shows a straight segment of a conducting
50
wire, with length 1 and cross-section area A; the current is
from bottom to top. The wire is in a uniform magnetic field B,
30

perpendicular to the plane of the diagram and directed into the


plane. Let us assume first that the moving charges are positive,
later we shall see what happens when they are negative.

-159-
Figure (10.7): Strait segment of conducting.
The drift velocity Vd is upward, perpendicular to B. The
force F on each charge is
F = q(vd x B),
directed to the left, as shown in the figure. In this case, v d and
B are perpendicular, and the magnitude F of the force is

0
F= qvd B.

10
Using the same Eqns.(6.2) and (6.3), an expression for

40
the total forces can be derived on all the moving charges in a
length  of conductor with cross-section area A. Consider the
51
number of charges per unit volume is n and the number of
charges in segment of conductor with length  is nA  . The
50
total force F on all the moving charges in this segment has
magnitude.
30

F = (n A  )(q vdB)
= (nqvd)(A  B). (10.12)

The current density J, from Eqn. (6.3), is J = nqv d, and


the product JA is the total current I, so Eqn. (10.12) can be
rewritten as:
F= I B (10.13)

-160-
If the B field is not perpendicular to the wire but makes
an angle  with it, we handle the situation the same way we
did for a single charge. The component of B parallel to the
wire (and to the drift velocities of charges) exerts no force; the
component perpendicular to the wire is B ⊥ = B sin . The
general relation is:

0
F= I  B⊥ = I  B sin (10.14)

10
The force is always perpendicular to both the conductor

40
and the field, with the direction determined by the same right-
hand rule that we used for a moving positive charge (Fig.10.7).
51
We see that this force be expressed as a vector product, just as
the force on a single moving charge can. The segment of wire
50
is represented with a vector  along the wire in the direction
30

of the current; then the force F on this segment is:


_ _
F = I x B (10.15)

Figure (10.8) illustrates the direction relation for several cases.


Note that I is not a vector; the direction of charge motion is
described by .

-161-
0
10
Figure (10.8): Direction relation.

40
If the conductor is not straight, we can divide it into
infinitesimal segments dl; the force d F on each segment is:
51
dF = I dl x B. (10.16)
50
30

Then we can integrate this wire to find the total force on


a conductor of any shape. The integral is a line integral, the
same mathematical operation that we have used to define work
and potential (section 4.2). Finally, what happens when the
moving charges are negative, such as electrons in a metal?
Then in Fig.(10.7) an upward current corresponds to a
downward drift velocity, but because q is negative in this case,
the direction of the force F is the same as before. Thus,

-162-
Eqn.(10.13) through (10.16) are valid for both positive and
negative charges and even when both signs of charge are
present at once. This happens in some semiconductor materials
and in ionic solutions.

10.5. Force Between Parallel Conductors:

0
10
The interaction force between two long, current-
carrying conductors is important in a variety of practical

40
problems, and it also has fundamental significance in
connection with the definition of the ampere. Figure (10.9)
51
shows segments of two long, straight, parallel conductors
separated by a distance r and carrying currents I and I’,
50
respectively, in the same direction. Each conductor lies in the
magnetic field set up by the other, so each experience a force.
30

The diagram shows some of the field lines set up by the


current in the lower conductor. From Eqn. (9.9) the magnitude
of the B vector at the upper conductor is:
oI
B=
2 r

-163-
0
10
Figure (10.9): Segments of two straight parallel

40
conductors.

51
From Eqn. (10.13) the force on a length L of the upper
conductor is:
 o II  L
50
F= I’L B = ,
2r
30

and the force per unit length F/L is :


 o II
F/L= (8-11)
2r

The right-hand rule shows that the direction of the force on the
upper conductor is downward.

-164-
The current in the upper conductor also sets up a field at
the position of the lower one. Two successive applications of
the right-hand rule show that the force on the lower conductor
is upward. Thus two parallel conductors carrying currents in
the same direction attract each other. If the direction of either
current is reversed, the forces also reverse. Parallel conductors

0
carrying currents in opposite directions repel each other.

10
The fact that two straight, parallel conductors exert

40
forces of attraction or repulsion on one another is the basis of
the official SI definition of the ampere. One ampere is that
51
unvarying current which, if present in each of two parallel
conductors of infinite length and one meter apart in empty
50
space, causes each conductor to experience a force of exactly 2
x 10-7 Newton’s per meter of length.
30

This is an operational definition; it gives us an actual


experimental procedure for measuring current and defining a
unit of current. In principle, we could use this definition to
calibrate an ammeter, using only a meter stick and a spring
balance. For high-precision standardization of the ampere,
coils of wire are used instead of straight wires, and their
separation is made only a few centimeters. The complete
-165-
instrument, which is capable of measuring currents with a high
degree of precision, is called a current balance. This definition
also forms the basis of the SI definition of the coulomb as the
amount of charge transferred in one second by a current of one
ampere.

0
Mutual forces of attraction exist not only between wires

10
carrying current in the same direction, but also between the
longitudinal elements of a single current-carrying conductor.

40
If the conductor is a liquid or an ionized gas (a plasma), these
forces result in a constriction of the conductor, as if its surface
51
were acted on by an external, inward pressure. The
constriction of the conductor is called the pinch effect. The
50
high temperature produced by the pinch effect in a plasma has
been used in one technique to bring about nuclear fusion.
30

-166-
10.6. Ampere’s Law:

Ampere’s law provides an alternative formulation of the


relation between a magnetic field and its sources. In situations
with a high degree of symmetry, it is sometimes easier to use
than the law of Biot and Savart. The role of ampere’s law for

0
magnetic field is analogous to that of Gauss’s law for electric

10
fields, which we studied in chapter 2. We found that we could
use Gauss’s law to find the electric field caused by a highly

40
symmetric charge distributions corresponding to particular
electric field configurations. Similarly, we can use ampere’s
51
law to find magnetic fields caused by some highly symmetric
current distributions and to find current distributions
50
corresponding to particular magnetic field configurations.
Ampere’s law is formulated in terms of the line integral
30

of B around a closed path, denoted by:

 B.dl .
This is the same sort of integral that we used to define electric
potential in chapter 3. To evaluate this integral, we divide the
path into infinitesimal segments dl, calculate the scalar product
B.dl for each segment, and sum these products. In general, B

-167-
varies from point to point, and we must use the field at the
location of each dl. An alternative notation is

 Bdl,
where B is the component of B parallel to dl at each point.
The circle on the integral sign indicates that this integral is

0
always computed for a closed path, one whose beginning and
end points are the same. To introduce the basic idea, let’s

10
consider again the magnetic field caused by a long, straight

40
conductor carrying a current I. In section 9.4 it is found that
the field at a distance r from the conductor has magnitude:
oI
B=
51
2 r
and that the magnetic field lines are circles centered on the
50
conductor. If the line integral of B is taken around a circle with
radius r (a field line), B is constant and equal to B at every
30

point on the circle. The line integral of B for this path is just
the magnitude B times the circumference of the circle:
oI
 B.dl = 2r ( 2r )
= oI

-168-
The line integral is independent of the radius of the
circle and is equal to o multiplied by the current passing
through the area bounded by the circle.

We can also derive this result for a more general


integration path, as the one in Fig. 8-11a. At the position of

0
the line element dl the angle between dl and B is , and:

10
B.dl = B dl cos 

40
51
50
30

Figure (10.11): Integration path.


From the figure, dl cos = r d, where d is the angle
subtended by dl at the position of the conductor and r is the
distance of dl from the conductor. Thus:

-169-
oI
 Bdl =  2 r
(rd )

 I
= o  d
2

0
But  d is just the total angle swept out by the radial

10
line from the conductor to dl during a complete trip around the
path, that is, 2, so we get:

40
 B.dl =  o I (10.18)
51
This result for the line integral does not depend on the
shape of the path or on the position of the wire inside it. If the
50
current in the wire is opposite to that shown, the integral has
the opposite sign. But if the path does not enclose the wire
30

(Fig.10.11b), then the net change in  during the trip around


the integration path is zero, and the line integral is zero.

Now suppose several long, straight conductors pass


through the surface bounded by the integration path. The total
magnetic field B at any point on the path is the vector sum of
the fields produced by the individual conductor. Thus the line

-170-
integral of the total field B equals o times the algebraic sum
of the currents. To define this sum unambiguously, we need a
sign rule for the currents. For the surface bounded by the
Ampere-law path, wrap the fingers of your right hand around a
line perpendicular to this surface so that your fingers curl
around in the same direction in which you plan to go around

0
the path when you evaluate  B.dl . Then, your thumb indicates

10
the positive current direction. Currents that pass through in this
direction are positive; those in the opposite direction are

40
negative. Here’s another way to say the same thing: Looking
at the surface, integrate counterclockwise around the
51
boundary, currents moving toward you through the surface are
positive, and those going away from you are negative.
50

If the integration path does not enclose a particular wire,


30

the line integral of the B field of that wire is zero because the
angle  for that wire sweeps through a net charge of zero
rather than 2 during the integration. Any conductors present
that are not enclosed by a particular path may still contribute to
the value of B at every point, but the line integrals of their
fields around the path are zero.

-171-
We conclude that the general case is obtained by
replacing, in Eqn. (10.18) with Iencl, the algebraic sum of the
currents enclosed or linked by the integration path, with the
sum evaluated using the sign rule just described. The general
form of Ampere’s law is:

 B.dl =  o I encl (10.19)

0
10
We have derived this relation only for the special case
of the field of several long, straight parallel conductors.

40
However Eqn. (10.19) is valid for conductors and paths of any
shape. The general derivation is no different in principle from
51
what we have presented, but the geometry is more
complicated.
50

When  B.dl = 0 , this does not necessarily mean that


30

B=0 everywhere along the path; it means only that the total
current through an area bounded by the path is zero. In the
form we have stated it, Ampere’s law is valid only for steady
currents and for magnetic fields that do not vary with time.

-172-
References:
1. F.W. Sears and M.W. Zemansky:
”University Physics”, Addison-Wesley Pub. Co.,
Inc.

0
2. D. Halliday and R. Resnick: “Physics”, John

10
Wiley & Sons, Inc.

40
51
50
30

-173-

You might also like