22 Scheme Physics for Eee Module 4 Notes

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 34

AMC ENGINEERING COLLEGE

DEPARTMENT OF PHYSICS

Module 4 Notes

Maxwell’s Equations and


Electromagnetic Waves

I/II SEMESTER Physics for EEE Stream


Subject code: BPHYE102/202
This page was intentionally left blank.
Contents

1 Maxwell’s Equations 5
1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.2 Fundamentals of Vector Algebra . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.2.1 Vector . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.2.2 Basis Vectors/Coordinate System . . . . . . . . . . . . . . . . . . . . . 6
1.2.3 Dot Product . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.2.4 Cross Product . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.3 Fundamentals of Vector Calculus . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.3.1 Scalar and Vector Fields . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.3.2 Del Operator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.3.3 Gradient . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.3.4 Divergence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.3.5 Curl . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.3.6 Worked Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
1.4 Types of Integrals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
1.4.1 Line Integral . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
1.4.2 Surface Integral . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
1.4.3 Volume Integral . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
1.5 Laws of Electric and Magnetic Fields . . . . . . . . . . . . . . . . . . . . . . . 18
1.5.1 Gauss’s Law for electric fields . . . . . . . . . . . . . . . . . . . . . . . 18
1.5.2 Gauss’s Law for magnetic fields . . . . . . . . . . . . . . . . . . . . . . 18
1.5.3 Faraday’s Law of Electromagnetic Induction . . . . . . . . . . . . . . . 18
1.5.4 Biot-Savart Law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
1.5.5 Ampere’s Law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
1.6 Gauss’s Divergence Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
1.7 Stokes’ Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
1.8 Differential/Point forms of the laws of electric and magnetic fields . . . . . . 22
1.9 The Continuity Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
1.10 Displacement Current . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
1.10.1 Expression for Displacement Current . . . . . . . . . . . . . . . . . . . 24
1.11 Maxwell’s Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
1.12 Maxwell’s Equations in Vacuum . . . . . . . . . . . . . . . . . . . . . . . . . . 26
1.13 Model and Previous Year Questions . . . . . . . . . . . . . . . . . . . . . . . 26
1.14 Numericals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26

2 Electromagnetic Waves 27
2.1 Wave Equation of Electric Field in Vacuum . . . . . . . . . . . . . . . . . . . 27
2.2 Plane Waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.3 Transverse nature of Electromagnetic Waves . . . . . . . . . . . . . . . . . . . 29

3
CONTENTS

2.4 Polarization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
2.4.1 Linear Polarization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
2.4.2 Circular Polarization . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
2.4.3 Elliptical Polarization . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
2.5 Model and Previous Year Questions . . . . . . . . . . . . . . . . . . . . . . . 34

Department of Physics 4 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
Chapter 1

Maxwell’s Equations

Syllabus
Maxwell’s Equations
Fundamentals of Vector Calculus. Divergence and Curl of Electric field and Magnetic
field (static), Gauss’ divergence theorem and Stoke’s theorem. Description of laws of
Electrostatics, Magnetism, Faraday’s laws of EMI, Current Density, Equation of Conti-
nuity, Displacement Current (with derivation), Maxwell’s equations in vacuum, Numer-
ical Problems

1.1 Introduction
In the study of engineering of electronics and allied branches, understanding the behavior of
electric and magnetic fields is crucial. These fields, which permeate space around charged
and magnetic objects, respectively, are vector fields. Vector calculus provides powerful tools
for analyzing and interpreting these fields.
This chapter delves into two fundamental concepts: divergence and curl. These concepts
are essential for characterizing the behavior of vector fields, particularly electric and mag-
netic fields in static situations, where the fields do not change with time. We will explore the
mathematical definitions of divergence and curl, their physical interpretations, and their ap-
plications in solving practical problems related to electric and magnetic fields. Understanding
these concepts will not only deepen your knowledge of vector calculus but also provide you
with a solid foundation for tackling more advanced topics in electromagnetism.

1.2 Fundamentals of Vector Algebra


1.2.1 Vector
A vector is a quantity which has both a magntiude and direc-
tion. In physical space, it is represented by an arrow, where the
length of the arrow represents the magnitude of the vector and
the direction of the arrow-head represents the direction of the
vector. Symbolically, a vector is represented by an alphabet
which is bold-faced, like b, or with an arrow symbol on top,
like ⃗b. We will use the latter form in this book. If we write
Figure 1.1: A vector
5
1.2. FUNDAMENTALS OF VECTOR ALGEBRA

an alphabet with no arrow on top, then it would be a scalar


quantity.
For example, consider a vector as shown in figure 1.1. Symbolically, this vector can be
written as
⃗b = bb̂

where ⃗b represents the vector, b is the magnitude of the vector, and b̂ is the direction of the
vector. From the figure, b tells us the length of the arrow, and b̂ tells us the direction in
which the arrow is pointing.
Note that here, b̂ is also a vector, but it is special in the sense that its magnitude is 1.
Such a vector whose magnitude is unity is called as a unit vector and is symbolically denoted
by a ‘carrot’ or a ‘hat’ symbol (ˆ) on top.

1.2.2 Basis Vectors/Coordinate System


If we want to describe vectors using numbers, instead of just drawing them as arrows, we
will need something called as a coordinate system. A coordinate system is formed by a set
of special vectors called as basis vectors. These vectors are special in the sense that any
arbitrary vector can be written as a combination of these vectors.
For example, one of the most popular coordinate systems in 3-dimensions, and the one we
will be using in this chapter, is the ‘3-dimensional Cartesian coordinate system’. This coordi-
nate system has 3 basis vectors which are denoted as î, ĵ, and k̂ or equivalently x̂, ŷ, and ẑ.
These denote the unit vectors in the X-, Y-, and Z-axes respectively. So given the basis
vectors, we will be able to describe any arbitrary vector using a set of numbers.

Figure 1.2: Vector Resolution

For example, consider some vector ⃗b as shown in figure 1.2. For simplicity, we will consider
here only 2 dimensions (î and ĵ), but the idea extends to 3 dimensions also. If we draw the
X- and Y-axis using the basis vectors, and draw perpendicular lines from the arrow tip to
both axes, we will see how much each basis vector contributes in creating the given vector. In
our case, there are 4 units of contribution from î and 2 units of contribution from ĵ. Hence
we can write the given vector ⃗b as a sum of the basis vectors as follows:

⃗b = 4î + 2ĵ

This is called as the resolution of a vector in a given coordinate system. The numbers 4
and 2 are called coefficients or coordinates of the vector in the given coordinate system. So
in general, an arbitrary vector ⃗a can be resolved in three-dimensions as follows:

⃗a = ax î + ay ĵ + az k̂

Department of Physics 6 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
1.2. FUNDAMENTALS OF VECTOR ALGEBRA

where, ax , ay , and az are the coordinates of ⃗a using respectively î, ĵ, and k̂ basis vectors,
i.e., the Cartesian coordinate system.
Sometimes, instead of writing the vector as a sum of basis vectors, it is common to write
only the coordinates, while ignoring the basis vectors. For example, our previous vector ⃗a
can be written simply as ⃗a = (ax , ay , az ). In such cases, the basis vectors are implied to be
the Cartesian basis vectors unless otherwise specified.
An alternate way is to write the coordinates as the elements of a column matrix like
 
a
 x
⃗a = ay 
 
 
az

1.2.3 Dot Product


The dot product is one way to “multiply” two vectors. The dot
product of two vectors ⃗a and ⃗b is defined as follows:

⃗a · ⃗b = ab cos(θ)

where a, b are the magnitudes of vectors ⃗a and ⃗b respectively,


and θ is the angle between the vectors when they are connected
tail-to-tail. Refer figure 1.3.
Figure 1.3: Dot product
Properties of the dot product
ˆ Notice that the RHS above is a scalar quantity. So the dot product of two vectors gives
a scalar and thus the dot product is also called the scalar product.

ˆ If the dot product of two vectors is 0, then the vectors are perpendicular to each other.

ˆ The dot product is commutative, i.e., ⃗a · ⃗b = ⃗b · ⃗a

Also remember the dot products between the Cartesian basis vectors (which you can
prove yourself):
î · î = 1 ĵ · ĵ = 1 k̂ · k̂ = 1
î · ĵ = 0 ĵ · k̂ = 0 k̂ · î = 0
In physics, the dot product is used, for example, to find out the work done on an object
⃗ acts on an object such that it moves a distance ⃗s, then the work
due to a force. If force F
done W is given by:
W =F ⃗ · ⃗s = F s cos(θ)

The dot product can also be written in terms of the vectors’ coordinates. So if ⃗a =
ax î + ay ĵ + az k̂ and ⃗b = bx î + by ĵ + bz k̂, then we have

⃗a · ⃗b = (ax î + ay ĵ + az k̂) · (bx î + by ĵ + bz k̂)


⃗a · ⃗b = ax bx + ay by + az bz

where we have opened the brackets, distributed everything, and simplified using the relations
of the dot products of basis vectors.
So in terms of the coordinates, the dot product of two vectors is the sum of the product
of the corresponding coordinates.

Department of Physics 7 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
1.3. FUNDAMENTALS OF VECTOR CALCULUS

1.2.4 Cross Product


The cross product is another way to “multiply” two vectors. The cross product of two vectors
⃗a and ⃗b is defined as follows:
⃗a × ⃗b = ab sin(θ)n̂
where n̂ is a unit vector pointing in a direction perpendicular to the plane containing ⃗a and
⃗b, and the other terms are the same as before. The unit vector n̂ is such that it follows
the right-hand screw rule: if the four fingers of the right hand are pointing in the direction
of ⃗a and are then curled towards the direction of ⃗b, then the direction of n̂ is given by the
direction in which the thumb is pointing.

Properties of the cross product


ˆ Notice that the RHS above is a vector quantity. So the cross product of two vectors
gives a vector and thus the cross product is also called the vector product.
ˆ The cross product is non-commutative, i.e., ⃗a × ⃗b ̸= ⃗b × ⃗a in general, but there is a
relation between the two: ⃗a × ⃗b = −⃗b × ⃗a.
Also remember the cross products between the Cartesian basis vectors (which you can
prove yourself):
î × î = ⃗0 ĵ × ĵ = ⃗0 k̂ × k̂ = ⃗0
î × ĵ = k̂ ĵ × k̂ = î k̂ × î = ĵ
ĵ × î = −k̂ k̂ × ĵ = −î î × k̂ = −ĵ
The cross product can also be written in terms of the vectors’ coordinates. So if ⃗a =
ax î + ay ĵ + az k̂ and ⃗b = bx î + by ĵ + bz k̂, then we have
⃗a × ⃗b = (ax î + ay ĵ + az k̂) × (bx î + by ĵ + bz k̂)
⃗a × ⃗b = (ay bz − az by )î + (az bx − ax bz )ĵ + (ax by − ay bx )k̂
where we have opened the brackets, distributed everything, and simplified using the relations
of the cross products of basis vectors.
If this formula looks confusing, there is a neat way of writing it in terms of a determinant
which looks like this:
î ĵ k̂
⃗a × ⃗b = a a ax y z

bx by bz
where the 1st row contains the Cartesian basis vectors, the 2nd row contains the coordinates
of the first vector and the 3rd row contains the coordinates of the 2nd vector. Calculating
the determinant yields the same formula as above, which you can verify.

1.3 Fundamentals of Vector Calculus


1.3.1 Scalar and Vector Fields
Scalar Fields
A scalar field is a function that associates a single number (scalar) to every point in a space.
These fields are used to describe physical quantities that have magnitude but no direction,
such as temperature, density, pressure, etc.

Department of Physics 8 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
1.3. FUNDAMENTALS OF VECTOR CALCULUS

A scalar field may be written mathematically as, for example, ϕ(x, y, z) = 3x2 y + 5xyz,
where we have considered the 3D Cartesian coordinate system. For our purpose in this
chapter, we will consider those scalar fields which take a 3D vector (x, y, z) as input and give
a scalar as the output.

Vector Fields
A vector field is a function that associates a single vector to every point in a space. These
fields are used to describe physical quantities that have magnitude and direction, such as
velocity, force, and electric or magnetic fields.
A vector field may be written mathematically as, for example, F ⃗ (x, y, z) = x2 y î − 2yz ĵ +
4xyz 2 k̂, where we have considered the 3D Cartesian coordinate system. Or in general, any
vector field can be written as
⃗ (⃗r ) = Fx (⃗r )î + Fy (⃗r )ĵ + Fz (⃗r )k̂
F

where ⃗r is called the position vector given by ⃗r = xî + y ĵ + z k̂ = (x, y, z) and Fx , Fy , Fz


are scalar fields and are each functions of (x, y, z). For our purpose in this chapter, we will
consider those vector fields which take a 3D position vector (x, y, z) as input and give a 3D
vector as the output.
Understanding and working with vector fields, in particular - electric fields and magnetic
fields - will be focus of this chapter. On that note, let me state here that there are mainly two
ways to represent each of these fields. Most of the time, the electric field can be represented
directly which is denoted by E. ⃗ But sometimes, it can be represented by something called
the ‘electric displacement’ which is denoted by D. ⃗ In fact, these two are related:

⃗ = ϵE
D ⃗

where ϵ is the absolute permittivity of the medium.


In this module, we will be only dealing with vacuum as a medium. So we can replace the
absolute permittivity with the permittivity of free space, ϵ0 :
⃗ = ϵ0 E
D ⃗ (1.1)

Next, the magnetic field is usually represented by the ‘magnetic flux density’ denoted by
⃗ But sometimes, it can be represented by something called the ‘magnetic field intensity’
B.
⃗ In fact, again, these two are related:
which is denoted by H.

⃗ = 1B
H ⃗
µ
where µ is the absolute permeability of the medium.
Again, we will be only dealing with vacuum as a medium in this module. So we can
replace the absolute permeability with the permeability of free space, µ0 :

⃗ = 1B
H ⃗ (1.2)
µ0

1.3.2 Del Operator


We know how to take derivatives of scalar functions. For example, if we have a simple 1D
scalar function f (x) = x2 , then its derivative with respect to x can be written as df /dx = 2x.
But if we have a 2D or 3D scalar function, then we have to make use of partial derivatives.

Department of Physics 9 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
1.3. FUNDAMENTALS OF VECTOR CALCULUS

For example, if g(x, y) = xy + y 2 , then the derivative of g with respect to only y will be
written as ∂g/∂y = x + 2y.
Now, what if I wanted to take the derivative of a 2D or 3D scalar function with respect
to all variables? This can be done by using something called as the del operator, which is
denoted by the symbol ∇, and is written as a “vector”:
∂ ∂ ∂
∇ = î + ĵ + k̂
∂x ∂y ∂z
where we have considered the 3D Cartesian coordinate system (mentioning this is important
because this operator will look different in different coordinate systems).
The del operator looks like a vector, but it is a vector operator. By itself, it is just
a collection of partial derivatives in the form of a vector and does not have any physical
meaning, unless it ‘operates’ on something. The del operator can operate on both scalar
fields and vector fields in different ways and each type has a different physical interpretation.
These operation types are called the gradient, divergence, and curl, which we will discuss
next.

1.3.3 Gradient
The gradient is an operation where the del operator operates on a scalar field and gives a
vector field. (A gradient can also operate on vector fields, but we will only consider scalar
fields here.) If ϕ(x, y, z) is a scalar field, then its gradient is written as ∇ϕ(x, y, z).
For example, if ϕ(x, y, z) = xyz + y 2 z, then the gradient of ϕ is
 
∂ ∂ ∂
∇ϕ = î + ĵ + k̂ ϕ
∂x ∂y ∂z
∂ϕ ∂ϕ ∂ϕ
= î + ĵ + k̂
∂x ∂y ∂z
∂(xyz + y z)2 ∂(xyz + y 2 z) ∂(xyz + y 2 z)
= î + ĵ + k̂
∂x ∂y ∂z
∇ϕ = (yz)î + (xz + 2yz)ĵ + (xy + y 2 )k̂

1.3.4 Divergence
The divergence is an operation where the del operator operates on a vector field and gives a
scalar field. It can be thought of as taking the “dot product” of the del operator with the
given vector field. So if F ⃗ (x, y, z) is a vector field, its divergence is written as ∇ · F
⃗ (x, y, z).
⃗ ⃗
So if F = Fx î + Fy ĵ + Fz k̂, then the divergence of F can be written as

⃗ = ∂Fx + ∂Fy + ∂Fz


∇·F
∂x ∂y ∂z

Physical significance of divergence


The physical significance of the divergence of a vector field is closely tied to the concept of
flux, which is the flow of a vector quantity through a surface. The divergence of a vector field
at a point indicates the net outward flux of field lines per unit volume at that point.
ˆ If the divergence at a point is positive, it indicates that the vector field is spreading out
from that point. This is analogous to a ‘source’ of flow, where the field lines diverge
outward, like the flow of water from a sprinkler (see figure 1.4a and imagine the sprinkler
is at the origin).

Department of Physics 10 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
1.3. FUNDAMENTALS OF VECTOR CALCULUS

(a) Positive Divergence (b) Negative Divergence (c) Zero Divergence

Figure 1.4: Divergence of some 2D vector fields

ˆ If the divergence at a point is negative, it indicates that the vector field is converging
toward that point. This is analogous to a ‘sink’ of flow, where the field lines converge
inward, like the flow of water into a drain (see figure 1.4b and imagine the drain is at
the origin).

ˆ If the divergence at a point is zero, it indicates that the vector field is neither spreading
out nor converging at that point. This could mean that the field is circulating around
the point, or that the flow is balanced, with as much flow entering as leaving the point
(see figure 1.4c). Additionally, a vector field whose divergence is zero at all points is
said be solenoidal.

Figure 1.5: Electric field of a positive charge at the origin

Figure 1.5 shows the three-dimensional electric field of a positive charge (shown as a small
sphere; imagine this to be point-sized) placed at the origin. Only a small sample of vectors

Department of Physics 11 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
1.3. FUNDAMENTALS OF VECTOR CALCULUS

are shown to avoid clutter. The length and direction of each vector denotes respectively the
strength of the electric field and the direction in which a positive charge would move if placed
at the tail of the vector. You will observe that all the vectors are pointing away from the
origin, which means the divergence of this field at the origin is positive. Also note the length
of the vectors: arrows near the charge are longer than those away from the charge, as it
should be from Coulomb’s law.

Divergence of Electric Field


The divergence of electric field is found to be as follows:

∇·E ⃗ = ρ (1.3)
ϵ0
where E⃗ represents the electric field, ρ is called the charge density and ϵ0 (= 8.852 ×
10−12 F/m) is the permittivity of free space.
Or in terms of the electric displacement D,⃗ using (1.1), we can write the above as

⃗ =ρ
∇·D (1.4)

The charge density is a scalar field which contains information about how much charge
is present per unit volume at each point in space. So at regions where there are positive
charges, the divergence will be positive, which means the electric field spreads out, and at
regions where there are negative charges, the divergence will be negative, which means the
electric field converges in.

Divergence of Magnetic Field


The divergence of magnetic field is found to be as follows:
⃗ =0
∇·B (1.5)

where B ⃗ represents the magnetic field.


This is saying that the divergence of magnetic field is zero everywhere in space, which
means magnetic fields are solenoidal. This implies there are no “sources” or “sinks” of
magnetic fields. In others words, the expression suggests that magnetic monopoles do not
exist (as far as we know).

1.3.5 Curl
The curl is an operation where the del operator operates on a vector field and gives another
vector field. It can be thought of as taking the “cross product” of the del operator with the
⃗ (x, y, z) is a vector field, its curl is written as ∇ × F
given vector field. So if F ⃗ (x, y, z).
If the coordinates of the vector field is given, then we can use a similar formula as the
cross product to find the curl, i.e., using a determinant. So if F ⃗ = Fx î + Fy ĵ + Fz k̂, then the
⃗ can be written as
curl of F
î ĵ k̂
∇×F ⃗ = ∂ ∂ ∂
∂x ∂y ∂z

Fx Fy Fz
     
⃗ = ∂Fz − ∂Fy î + ∂Fx − ∂Fz ĵ + ∂Fy − ∂Fx k̂
=⇒ ∇ × F
∂y ∂z ∂z ∂x ∂x ∂y

Department of Physics 12 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
1.3. FUNDAMENTALS OF VECTOR CALCULUS

Figure 1.6: Examples of 2D vector fields with non-zero curl at certain points

Physical significance of curl


The physical significance of the curl of a vector field is related to the concept of circulating
behavior of the field around a point. The curl of a vector field at a point indicates the
tendency of the field to circulate per unit area at that point.

ˆ If the curl is non-zero at a point, it indicates the field lines form closed loops around
the point, indicating the presence of a circulating motion. If the curl is zero at a point,
it means that the field lines do not form closed loops around the point. A vector field
whose curl is zero at all points is said be irrotational.

ˆ The magnitude of the curl at a point gives the strength of the rotational behavior of the
field around that point. A larger magnitude indicates a stronger rotational component,
while a smaller magnitude indicates a weaker rotational component.

ˆ The direction of the curl at a point gives the axis of rotation of the field around that
point, according to the right-hand rule. For example, if the thumb of the right hand is
pointing in the direction of the curl, then the way in which the four fingers curl around
the thumb will be the way in which the vector field curls around that point.

Figure 1.7 shows the three-dimensional magnetic field of a infintely long wire (shown as
a thin cylinder along the Z axis) caryying current in the positive Z direction. Only a small
sample of vectors are shown to avoid clutter. The length and direction of each vector denotes
respectively the strength and direction of the magnetic field at the tail of the vector. You will
observe that the vectors are curling around the Z-axis in the positive sense (counter-clockwise
direction or from positive-X axis towards positive-Y axis), which means the curl of this field
is pointing along the positive Z direction. Also note the length of the vectors: arrows near
the wire are longer than those away from the wire, as it should be from the Biot-Savart’s
law.

Curl of Electric Field (static case)


A field is said to be static if it does not change with time. The curl of a static electric field
is found to be as follows:
∇×E ⃗ =0 (1.6)
where E ⃗ represents the electric field.
This is saying that the curl of a static electric field is zero, which means static electric
fields are irrotational.

Department of Physics 13 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
1.3. FUNDAMENTALS OF VECTOR CALCULUS

Figure 1.7: Magnetic field of an infinite wire carrying current in the +Z direction

Curl of Magnetic Field (static case)


The curl of a static magnetic field is found to be as follows:
⃗ = µ0 J
∇×B ⃗ (1.7)
⃗ represents the magnetic field, J
where B ⃗ is called the current density, and µ0 (= 4π ×
−7
10 H/m) is the permeability in free space.
⃗ using (1.2), we can write the above as
Or in terms of the magnetic field intensity H,
⃗ =J
∇×H ⃗ (1.8)

The current density is a vector field which contains information about how much current
is flowing per unit area (imaginary infinitesimal patch placed perpendicular to the direction
of current) at each point in space. The direction of the current density is, by convention, the
direction in which positive charge flows. For example, if the current density is in the positive
Z direction, then the magnetic field due to this current curls around from the positive X
direction towards the positive Y direction just like in figure 1.7.

1.3.6 Worked Examples


⃗ = 3x2 î + 1 y 2 z ĵ + 3xy k̂ at point (1, 2, 4).
Q 1. Find the divergence of the vector field A 2
Given:
⃗ = 3x2 î + 1 y 2 z ĵ + 3xy k̂
A
2
Point (x, y, z) = (1, 2, 4)
⃗ =?
To find: ∇ · A
   
⃗ = ∂ ∂ ∂ 2 1 2
∇·A î + ĵ + k̂ · 3x î + y z ĵ + 3xy k̂
∂x ∂y ∂z 2

Department of Physics 14 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
1.4. TYPES OF INTEGRALS

⃗ = ∂(3x2 ) ∂( 12 y 2 z) ∂(3xy)
∇·A + +
∂x ∂y ∂z
⃗ = 6x + yz
=⇒ ∇ · A
Then, at the given point, the divergence will be

⃗ (1,2,4) = 6(1) + (2)(4) = 14


∇·A

⃗ = (x + ay)î + (y + bz)ĵ + (x + cz)k̂ is solenoidal.


Q 2. Find constant c such that A
Given:
⃗ = (x + ay)î + (y + bz)ĵ + (x + cz)k̂
A
⃗ = 0 since it is solenoidal
∇·A
To find: c =?

⃗ =0
∇·A

∂(x + ay) ∂(y + bz) ∂(x + cz)


+ + =0
∂x ∂y ∂z
1+1+c=0
=⇒ c = −2

⃗ = (1 + yz 2 )î + xy 2 ĵ + x2 y k̂.
Q 3. Calculate the curl of the vector field A
Given:
⃗ = (1 + yz 2 )î + xy 2 ĵ + x2 y k̂
A
⃗ =?
To find: ∇ × A

î ĵ k̂
⃗ =
∇×A ∂ ∂ ∂
∂x ∂y ∂z

(1 + yz 2 ) xy 2 x2 y

∂(x2 y) ∂(xy 2 ) ∂(1 + yz 2 ) ∂(x2 y) ∂(xy 2 ) ∂(1 + yz 2 )


     
⃗ =
∇×A − î + − ĵ + − k̂
∂y ∂z ∂z ∂x ∂x ∂y

⃗ = x2 î + (2yz − 2xy)ĵ + (y 2 − z 2 )k̂


∇×A

1.4 Types of Integrals


1.4.1 Line Integral
A line integral is an expression of the form
Z Z ⃗b
⃗ · d⃗l =
A ⃗ · d⃗l
A
L ⃗
a

Department of Physics 15 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
1.4. TYPES OF INTEGRALS

Figure 1.8: 2D Vector field with a curve

where A ⃗ is a vector field, d⃗l is an infinitesimal displacement vector, and the integral is carried
out along a given path L from ⃗a to ⃗b (see figure 1.8, where we have taken a 2D vector field
and a 2D curve, but the idea extends to 3D as well).
The line integral calculates how closely the vector field aligns with the given curve. To
do this, we consider a small line element d⃗l along the curve. For example, from the figure,
consider point M and draw a tangent to the curve at this point (green arrow in the figure;
it is scaled up for clarity, but think of it to be infinitesimally small). This tangent is our d⃗l.
Next, find the dot product of the vector (of the field) present at this point and the tangent
vector, i.e., A⃗ · d⃗l. This dot product tells us how closely aligned the tangent line of the curve
is to the vector from the field at this point. But we want to find this alignment for all the
points on the curve, so we integrate the term A ⃗ · d⃗l from the start to the end of the curve,

i.e., from ⃗a to b, which gives the above expression.
If the path in question forms a closed loop, i.e., ⃗b = ⃗a, then we write the line integral as
I
⃗ · d⃗l
A
L

where the circle on the integral symbol denotes a closed loop. This particular integral is also
called the circulation of the vector field around the path L.

1.4.2 Surface Integral


A surface integral is an expression of the form
Z
⃗ · d⃗s
A
S

where A⃗ is a vector field, d⃗s is an infinitesimal patch of area with a direction perpendicular
to the surface, and the integral is carried out over the given surface S (see figure 1.9).
The surface integral calculates the flux of a vector field flowing through a surface. This
is done by considering how perpendicular the vector field is to the given surface. To do this,
we consider a small patch of area d⃗s on the surface. For example, from the figure, consider
the green patch having area ds. Draw a unit vector normal (or perpendicular) to the patch

Department of Physics 16 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
1.4. TYPES OF INTEGRALS

Figure 1.9: 3D vector field with a surface

at this point (n̂ in the figure). This unit vector gives the ‘direction’ of our small area. So the
area vector associated with the small green patch is d⃗s (= ds n̂). Next, find the dot product
of the vector (of the field) present at this point and the area vector, i.e., A ⃗ · d⃗s. This dot
product tells us how closely aligned the normal vector of our small area is to the vector from
the field at this point; in other words, this tells us how perpendicular the field’s vector is to
the surface at this point. But we want to find this perpendicularity for all the patches of
areas on the surface, so we integrate the term A ⃗ · d⃗s over the given surface, which gives the
above expression. When calculating the integration over a surface, we have to use a double
integral where the limits will be given by the boundary of the surface.
If the surface in question is closed, like the surface of a sphere, etc., then we write the
surface integral as I
⃗ · d⃗s
A
S

where the circle on the integral symbol denotes the surface S is closed.

1.4.3 Volume Integral


A volume integral is an expression of the form
Z
ϕ dτ
V

where ϕ is a scalar field, dτ is an infinitesimal volume element, and the integral is carried
out over the given volume V .
For example, if ϕ represents the mass density of a material (which might vary from point
to point), then the volume integral over the bulk of the material would give us the total mass
of the material. Or if ϕ represents the charge density in a region (which might vary from

Department of Physics 17 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
1.5. LAWS OF ELECTRIC AND MAGNETIC FIELDS

point to point), then the volume integral over the volume of this region would give us the
total charge in the region.
In Cartesian coordinates, the infinitesimal volume element is written as dτ = dx dy dz.
And of course, when calculating the integration over a volume, we have to use a triple integral
where the limits will be given by the closed surface bounding the volume.

1.5 Laws of Electric and Magnetic Fields


1.5.1 Gauss’s Law for electric fields
Gauss’s law provides us a way to calculate electric fields for charge distributions which show
symmetry.
Consider a small region in space consisting of charges. To find out the electric field at
any point in space due to this collection of charges, we will consider a closed surface which
encloses these charges. This imaginary surface is called a Gaussian surface S. If Q denotes
the total charge present within this surface, then Gauss’s law states that
I
⃗ · d⃗s = Q
E (1.9)
S ϵ0
where we have a closed surface integral of the electric field over the Gaussian surface and ϵ0
is the permittivity in free space.
⃗ we get
Or in terms of the electric displacement D,
I
D⃗ · d⃗s = Q (1.10)
S

Remember that the surface integral of a vector field represents the field’s flux (number
of field lines passing through a given area). So this law simply states that the flux of electric
field through a Gaussian surface is proportional to the charge enclosed within it.
So given Q and choosing a suitable Gaussian surface, we can find the electric field at any
point due to the charges.

1.5.2 Gauss’s Law for magnetic fields


The Gauss’s law for magnetic fields states that
I
B⃗ · d⃗s = 0 (1.11)
S

This means that the net flux of magnetic field through a closed surface is always zero.
This implies that the number of field lines entering and leaving the closed surface are the
same.

1.5.3 Faraday’s Law of Electromagnetic Induction


Faraday’s law states that a changing magnetic flux induces an EMF (electromotive force or
voltage) in a circuit linked to it. The induced EMF is proportional to the rate of change of
flux and opposes the change. This can be mathematically written as
dϕm
ε=−
dt
where ε is the induced EMF, ϕm is the magnetic flux and t is time.

Department of Physics 18 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
1.5. LAWS OF ELECTRIC AND MAGNETIC FIELDS

⃗ using the expression


The magnetic flux ϕm can be written in terms of the magnetic field B
Z
ϕm = ⃗ · d⃗s
B
S

where S is the surface bounded by the circuit loop and d⃗s is an infinitesimal patch of this
surface. Substituting this in the expression for EMF ε, we get
Z
d ⃗ · d⃗s
ε=− B
dt S
Z ⃗
∂B
ε=− · d⃗s
S ∂t

where taking the ordinary derivative inside the integral turns it into a partial derivative.
But the EMF is, by definition, the circulation of the electric field E⃗ around the circuit
loop. This can be mathematically written as
I
ε= ⃗ · d⃗l
E
L

where L represents the circuit loop and d⃗l is a line element of this loop.
So finally, Faraday’s law can be written in terms of only electric and magnetic field as
follows I Z
∂B⃗
⃗ · d⃗l = −
E · d⃗s (1.12)
L S ∂t
This expressions simply tells us how a time-varying magnetic field gives rise to an electric
field.

1.5.4 Biot-Savart Law

Figure 1.10: Biot-Savart Law

The Biot-Savart law is used to find the magnetic field at any point given some steady
(non-changing) current distribution. Consider a portion of a current carrying wire as shown
in figure 1.10. It carries current I in the specified direction. We want to find the magnetic
field produced at point P due to this current carrying wire. To do this, let us consider an

Department of Physics 19 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
1.5. LAWS OF ELECTRIC AND MAGNETIC FIELDS

infinitesimal line element d⃗l along the wire as shown. The vector drawn from the center of
this line element to point P is ⃗r , and the angle made by the line element with this vector is
θ. The Biot-Savart law states that the magnetic field dB ⃗ at point P due to this line element

dl is

ˆ directly proportional to current in the wire, I

ˆ directly proportional to the length of the line element, dl

ˆ directly proportional to sin(θ)

ˆ inversely proportional to the square of the distance from the line element to the point
P , i.e., r2

ˆ directed along n̂, which is the direction of d⃗l × r̂

Combining all of these together, we get

⃗ ∝ I dl sin(θ)
dB n̂
r2
⃗ = µ0 I dl sin(θ) n̂
dB
4π r2

⃗ = µ0 I dl × r̂
dB
4π r2

where µ0 /4π is the constant of proportionality and d⃗l × r̂ = dl sin(θ)n̂ is by defintion. µ0 is


called the permeability in free space.
⃗ we get
Or in terms of the magnetic field intensity H,

⃗ = 1 I d⃗l × r̂
dH
4π r2
For the configuration shown above, the magnetic field is directed into the plane of the
paper as shown by the ⊗ symbol. To find the total magnetic field at point P , we have to
integrate the above expression over the full length of the wire.

1.5.5 Ampere’s Law


The Ampere’s law provides us a way to calculate magnetic fields for some steady (non-
changing) current distributions which show symmetry.
Consider a long straight wire carrying current. To find out the magnetic field at any
point in space due to this wire, we will consider a closed loop which surrounds this wire.
This imaginary loop is called an Amperean loop. If I denotes the total current that passes
through the area enclosed by this loop - let’s call it L - then Ampere’s law states that
I
⃗ · d⃗l = µ0 I
B (1.13)
L

where we have a closed line integral of the magnetic field around the Amperean loop and µ0
is the permeability in free space.
⃗ we get
Or in terms of the magnetic field intensity H,
I
H⃗ · d⃗l = I (1.14)
L

Department of Physics 20 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
1.6. GAUSS’S DIVERGENCE THEOREM

Remember that the closed line integral of a vector field represents the field’s circulation.
So this law simply states that the circulation of magnetic field around an Amperean loop is
proportional to the current enclosed by it.
Note that this law is true for only steady currents, which means the magnetic field will
be static. For non-static magnetic fields, there is a small modification to this law which we
shall see later.

1.6 Gauss’s Divergence Theorem


⃗ , the Gauss’s divergence theorem is written as follows:
For a vector field F
Z I
⃗ ) dτ =
(∇ · F ⃗ · d⃗s
F
V S

It states that the integral of the divergence of a vector field over a volume is equal to the
field’s flux through the closed surface bounding the volume. Remember that the flux of a
vector field (RHS of above expression) is the surface integral of the normal component of a
vector field over a surface. In simple words, this theorem gives a relation between the volume
integral and the (closed) surface integral of a vector field.

Proof

We will try to prove this using Gauss’s law for electric fields. Consider a small region in space
filled with charges and let the density of these charges be denoted by ρ. Consider a Gaussian
surface S bounding these charges. Then by Gauss’s law (1.10), we have
I
⃗ · d⃗s = Q
D
S

where Q is the total charge enclosed within the Gaussian surface. But we can write the total
charge Q as a volume integral of the charge density ρ as follows
Z
Q= ρ dτ
V

where V is the volume bounded by the Gaussian surface.


Substituting this in the above expression, we get
I Z
⃗ · d⃗s =
D ρ dτ
S V

But we know from (1.4) that the integrand on the RHS is the divergence of the electric
⃗ = ρ. So we can finally write the above expression as
displacement i.e., ∇ · D
Z I
⃗ dτ =
(∇ · D) ⃗ · d⃗s
D
V S

which is the Gauss’s divergence theorem we wanted to prove.

Department of Physics 21 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
1.7. STOKES’ THEOREM

1.7 Stokes’ Theorem


⃗ , the Stokes’ theorem is written as follows
For a vector field F
Z I

(∇ × F ) · d⃗s = ⃗ · d⃗l
F
S L

It states that the integral of the normal component of the curl of a vector field over a
surface is equal to the circulation of the field around the loop bounding the surface. Remem-
ber that the circulation of a vector field (RHS of above expression) is the line integral of the
tangential component of the vector field around a closed loop. In simple words, this theorem
gives a relation between the surface integral and the (closed) line integral of a vector field.

1.8 Differential/Point forms of the laws of electric and mag-


netic fields
The four laws that we have discussed above, viz, Gauss’s law of electric field (1.9) (or (1.10))
and magnetic field (1.11), Faraday’s law (1.12), and Ampere’s law (1.13) (or (1.14)) can all
be written in differential form (also called point form) using the divergence theorem and the
Stokes theorem (which you can try to do yourself). On doing so, we get the following
ρ
⃗ = ⃗ =ρ
∇·E (or) ∇ · D (Gauss’s law for electric field) (1.15)
ϵ0
⃗ =0
∇·B (Gauss’s law for magnetic field) (1.16)

⃗ = − ∂B
∇×E (Faraday’s law) (1.17)
∂t
⃗ = µ0 J
∇×B ⃗ (or) ∇ × H ⃗ =J
⃗ (Ampere’s law) (1.18)

1.9 The Continuity Equation


Charge is a conserved quantity. This means that one cannot create or destroy charges; it
can move from one place to another. The mathematical form of this statement is called the
Continuity equation. To derive this equation, consider a region in space consisting of charges
moving about. Let us then consider an imaginary closed surface S somewhere in this region.
One can think of charges entering and leaving this imaginary surface. If J ⃗ is the current
flowing per unit area (also called current density) in our region, then the current I flowing
in and out of our closed surface can be written as a surface integral
I
I= ⃗ · d⃗a
J
S

But current I can also be written as the rate of change of charge. If we consider current
moving out of S to be positive, which means charges present inside S decreases, that means
we can write
dq
I=−
dt
where q is the charge present within the volume V bounded by the surface S. But we can
write this charge as a volume integral of the charge density over V . So we can write the

Department of Physics 22 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
1.10. DISPLACEMENT CURRENT

above expression as
Z
d
I=− ρ dτ
dt V

Z
∂ρ
I=− dτ
V ∂t

where taking the ordinary derivative inside the integral turns it into a partial derivative.
Comparing this with the surface integral above, we get
I Z
⃗ · d⃗a = − ∂ρ
J dτ
S V ∂t

Using the Gauss’s divergence theorem on the LHS, we get


Z Z
⃗ dτ = − ∂ρ
∇·J dτ
V V ∂t

And since this applies to any volume V , we can equate the integrands to finally get the
continuity equation
∇·J ⃗ = − ∂ρ (1.19)
∂t
In steady currents (or DC), which lead to static magnetic fields, the charge density has
to be unchanging with time, which means in such cases the continuity equation reduces to
∇·J ⃗ = 0. But in general, for non-steady currents (or AC), the divergence of current density
is non-zero.

1.10 Displacement Current


Equation (1.18), which is Ampere’s law, works well for steady currents but breaks down for
non-steady currents. To see this, take the divergence of (1.18) on both sides:

⃗ =∇·J
∇ · (∇ × H) ⃗

Here, the LHS is the divergence of a curl, which is always zero. But the RHS is generally
not zero for non-steady currents, but it should be zero. So Ampere’s law needs a modification
to take into account non-steady currents. This modification was done by Maxwell using the
continuity equation and it goes as follows:
Take the time derivative of (1.15)

∂ ⃗
 ∂ρ
∇·D =
∂t ! ∂t
∂D⃗ ∂ρ
∇· =
∂t ∂t

where we interchanged the del operator and time derivative operator on the LHS.

Department of Physics 23 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
1.10. DISPLACEMENT CURRENT

Using the continuity equation (1.19) on the RHS, we get


!
∂D⃗
∇· ⃗
= −∇ · J
∂t
!
∂D⃗
∇· +∇·J ⃗ =0
∂t
!
∂ ⃗
D
∇· J ⃗ + =0
∂t

The trick now is to use the term in brackets as the new current density in (1.18),


⃗ + ∂D
⃗ =J
∇×H
∂t
This is called the Maxwell-Ampere law and it works for both steady and non-steady
⃗ and is called the displacement
currents. The second term on the RHS has the same units as J
current density and is denoted as J⃗d , i.e.,

∂D⃗
J⃗d =
∂t
One should note that this displacement current density does not correspond to movement
of charges; it just has dimenions of current density. The name ‘displacement’ for J⃗d is given
due to its relation with the electric displacement D.⃗ But the name for D ⃗ was given in an
older theory of electromagnetism which has now been proven to be inaccuarte, but the name
has stuck.
The above are in terms of the electric displacement D ⃗ and magnetic field intensity H,.

⃗ ⃗
But they can also be written in terms of E and B as follows:


⃗ + µ0 ϵ0 ∂ E
⃗ = µ0 J
∇×B
∂t

∂E⃗
J⃗d = ϵ0
∂t
The presence of the displacement current density in Maxwell-Ampere law implies that a
time-varying electric field induces a magnetic field.

1.10.1 Expression for Displacement Current


The displacement current is the ‘current’ associated with a changing electric field; again, it
is not movement of charges. It is denoted by Id and is related to the displacement current
density through a surface integral Z
Id = J⃗d · d⃗s
S
where S is a surface through which we want to find the ‘current.’
We will try to get an expression for displacement current between the plates of a capacitor.
Consider a capacitor connected to an AC source as shown in figure 1.11.
Let A be the area of the capacitor plate and d be the distance between the plates. Consider
an external AC voltage V = V0 ejωt , where V0 is a constant, ω contains information about

Department of Physics 24 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
1.11. MAXWELL’S EQUATIONS

Figure 1.11: Capacitor connected to AC source


the frequency of the input signal and j is the imaginary number −1. (This is a common
way of writing voltage signals in electronics.)
The expression for displacement current is as given above
Z
Id = J⃗d · d⃗s
S

Let us consider the surface S to be inbetween and parallel to the plates and having the
same area A as the plates. Using this, we get J⃗d · d⃗s = Jd ds because both vectors point in
the same direction. Then considering Jd to be a constant through the surface, we can take
the integral only over ds which yields the area of our surface, which is A. So we get the
following expression
Id = J d A
But Jd = ∂D/∂t . So we have
∂D
Id = A
∂t
Let us write D in terms of the quantities we know. First, we know that D = ϵ0 E, where
E is the magnitude of electric field. Next, we know that the electric field can be written in
terms of voltage V and the distance between the plates d as E = V /d. Substituting these
above, we get
ϵ0 ∂V
Id = A
d ∂t
But we have V = V0 ejωt . So taking the time derivative of this, we have

∂V
= jωV0 ejωt
∂t
Substituting this in the displacement current expression, we finally get

ϵ0 ωV0 A jωt
Id = j e
d

1.11 Maxwell’s Equations


All the laws that we have studied until this point can be summarized into four equations
known as Maxwell’s equations. They are

Department of Physics 25 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
1.12. MAXWELL’S EQUATIONS IN VACUUM

ρ
⃗ = ⃗ =ρ
∇·E (or) ∇ · D (Gauss’s law for electric field)
ϵ0
⃗ =0
∇·B (Gauss’s law for magnetic field)

⃗ = − ∂B
∇×E (Faraday’s law)
∂t
⃗ ⃗
∇×B ⃗ + µ0 ϵ0 ∂ E (or) ∇ × H
⃗ = µ0 J ⃗ + ∂D
⃗ =J (Maxwell-Ampere law)
∂t ∂t
 
These equations along with the force law, F ⃗ =q E ⃗ + ⃗v × B
⃗ , summarize almost the
entire theoretical content of classical electrodynamics. Maxwell’s equations tell you how
charges produce fields; reciprocally, the force law tells you how fields affect charges.

1.12 Maxwell’s Equations in Vacuum


The Maxwell’s equations can also be written in the case where there are no charges or currents
⃗ = 0. In such cases, we get the Maxwell’s equations in vacuum:
in space, i.e., ρ = 0 and J
⃗ = 0 (or) ∇ · D
∇·E ⃗ =0 (1.20)
∇·B⃗ =0 (1.21)

∇×E⃗ = − ∂B (1.22)
∂t
⃗ ⃗
∇×B⃗ = µ0 ϵ0 ∂ E (or) ∇ × H
⃗ = ∂D (1.23)
∂t ∂t

1.13 Model and Previous Year Questions


Q 1. State and derive Gauss Divergence theorem. Mention Stoke’s theorem in mathematical
form.
Q 2. What is displacement current? Derive the expression for displacement current.
Q 3. State Biot-Savart’s law and list four Maxwell’s equations in differential form.
Q 4. Explain Faraday’s Laws of Electromagnetic induction, Amperes Law and express the
same in point form.
Q 5. What is vector opeartor ∇ ? Explain the concept of divergence, gradient, and curl.
Q 6. Discuss continuity equation. Derive the expression for displacement current.

1.14 Numericals
Q 1. Find the divergence of the vector field A ⃗ = 6x2 î + 3xy 2 ĵ + xyz 3 k̂ at point (1, 3, 6).
⃗ = 5y 4 z 2 î − 4x3 z 2 ĵ + 3x2 y 2 k̂ is solenoidal.
Q 2. Show that the vector field A
Q 3. Show that the vector field A ⃗ = e−ay (a cos(x)î − sin(y)ĵ) is solenoidal.
Q 4. If the electric displacement is D⃗ = 5x2 î − 3y 3 ĵ + 2z k̂, find the charge density at point
(2, -1, 3).
Q 5. For what values of a, b, and c, is the vector field A ⃗ = (3x2 + y + az)î + (bx − 5y 3 −
2
2z)ĵ + (2x + cy + 3z )k̂ irrotational?

Department of Physics 26 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
Chapter 2

Electromagnetic Waves

Syllabus
Electromagnetic Waves
The wave equation in differential form in free space (Derivation of the equation using
Maxwell’s equations), Plane Electromagnetic Waves in vacuum, their transverse nature.

2.1 Wave Equation of Electric Field in Vacuum


Equations (1.20) to (1.23) constitute a set of coupled, first-order, partial differential equations,
where E ⃗ and B
⃗ are related to each other. We can decouple them and write them as second-
order partial differential equations containing only E ⃗ or B ⃗ by themselves. Let’s try to do

this for E.
Consider equation (1.22)

∇×E ⃗ = − ∂B
∂t
Apply curl on both sides
!
  ⃗
∂B
⃗ = −∇ ×
∇× ∇×E
∂t

On the LHS, we need to use one of the del operator identities (which you can try to prove
yourself),
   
∇× ∇×A ⃗ =∇ ∇·A ⃗ − ∇2 A ⃗

where ∇2 is called the Laplacian operator and is simply ‘the dot product of the del operator
with itself’: ∇2 = ∇ · ∇
So we get
!
  ∂ ⃗
B
∇ ∇·E ⃗ − ∇2 E ⃗ = −∇ ×
∂t

Now, on the RHS, let us exchange the del and the time derivative operators

⃗ =− ∂ ∇×B
   
⃗ − ∇2 E
∇ ∇·E ⃗
∂t

27
2.2. PLANE WAVES

We can substitute equations (1.20) and (1.23) here


!

⃗ = − ∂ µ0 ϵ0 ∂ E
∇(0) − ∇2 E
∂t ∂t

Simplifying, we finally get


∂2E⃗
⃗ = µ0 ϵ0
∇2 E
∂t2
This is called as the wave equation of electric field in vacuum. Using a similar procedure,
you can also find the wave equation of magnetic field in vacuum:

∂2B
⃗ = µ0 ϵ0
∇2 B
∂t2
It is called a ‘wave equation’ because it looks like a well known equation, called the general
wave equation, from the theory of waves:
1 ∂ 2 u(x, t)
∇2 u(x, t) =
v 2 ∂t2
where u(x, t) represents some kind of wave, such as waves on a string, sound waves etc., and
v is the velocity of the wave. Comparing this equation with the electric field (or magnetic
field) wave equation, we see that
1
= µ0 ϵ0
v2
1
v=√
µ0 ϵ0
1
=√
4π × 10−7 × 8.852 × 10−12
= 2.998 × 108
v ≈ 3 × 108 = c

This means that electric and magnetic fields travel in vacuum with the speed of light.
The implication is astounding: perhaps light is an electromagnetic wave.

2.2 Plane Waves


The (3D) waves which vary only in the direction of its propagation, but remains uniform in
the other two perpendicular directions are called as plane waves. Figure 2.1 shows an example
of a plane wave propagating in the positive-Y direction. Only a small sample of vectors are
shown to avoid clutter. You will observe that the lengths and directions of the vectors are
different at different planes along the Y-axis. But on a particular plane, all vectors look the
same along both the X-axis and Z-axis.
Now, let’s write the mathematical expression for a plane wave. Consider a plane wave
propagating in the positive-Y direction. So the wave should be a function of only position y
and time t. Then we can write the plane wave as follows:
⃗ (y, t) = F⃗0 sin(ky − ωt + δ)
F

where F⃗0 is called the amplitude vector, δ is called the phase constant (which lies in the
range 0 ≤ δ < 2π), k is the wave number, and ω is the angular frequency of the wave.

Department of Physics 28 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
2.3. TRANSVERSE NATURE OF ELECTROMAGNETIC WAVES

Figure 2.1: Example of Plane Wave

The magntiude and direction of the amplitude vector gives the maximum amplitude and the
direction of oscillation of the wave respectively. The phase constant tells us how much the
wave has shifted along the direction of propagation. Also, k = 2π/λ and ω = 2πν, where λ
and ν are the wavelength and frequency of the wave respectively. Figure 2.1 is constructed
using this same expression with F⃗0 = 1k̂, δ = 0, λ = 8, ν = 0.75 in some appropriate units.
As it turns out, plane waves are one of the solutions to the general wave equation. This
means that we can write the electric field also as a plane wave (because it has a wave equation
associated with it). If we consider this electric field to be propagating along, say, the positive-
Y direction, we can write:

E(y, t) = E⃗0 sin(ky − ωt + δE )
where E⃗0 is the electric field amplitude vector and δE is the phase constant of the electric
field.
Similarly, we can write the magnetic field also as a plane wave (as it also has a wave
equation associated with it):

B(y, ⃗0 sin(ky − ωt + δB )
t) = B

where B⃗0 is the magnetic field amplitude vector and δB is the phase constant of the magnetic
field.

2.3 Transverse nature of Electromagnetic Waves


Now, we get the above expressions for electric and magnetic field plane waves from the theory
of waves. But these fields should also satisfy the Maxwell’s equations in vacuum. As the
Maxwell’s equations relate the electric field to the magnetic field, the above expressions for
the fields get greatly simplified and interesting conclusions can be drawn.
In particular, if we apply equation (1.20) and (1.21) to the above fields, we get the
following results

E⃗0 · ĵ = 0
B⃗0 · ĵ = 0

where ĵ is the direction of propagation in our example. This says that the electric and
magnetic fields are each perpendicular to the direction of propagation (remember that if the

Department of Physics 29 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
2.4. POLARIZATION

dot product of two vectors is zero, then they are perpendicular to each other), which means
that the direction of oscillation is perpendicular to the direction of propagation. These kinds
of waves are called as Transverse waves. (There is another class of waves, called longitudinal
waves, where the direction of oscillation is parallel to the direction of propagation. Sound
wave is an example.)
Next, if we apply (1.22) to the above fields, we get the following results

1. δE = δB
⃗0 = 1 ĵ × E⃗0
 
2. B (2.1)
c
where c is the speed of light in vacuum. These results state that
1. the electric field and magnetic fields are in phase.
⃗ is perpendicular to both E
2. B ⃗ and the direction of propagation. In other words, the
electric field, the magnetic field, and the direction of propagation are all mutually
perpendicular.
This is called as an electromagnetic wave (EM wave) in vacuum.
Let’s consider an example where δE = 0 and E⃗0 = E0 k̂, i.e., the electric field oscillates
along the Z-axis with amplitude E0 . Then the electric field plane wave can be written as

E(y, t) = E0 sin(ky − ωt)k̂

Equation (2.1) will become

⃗0 = 1 ĵ × E0 k̂
 
B
c
E 
⃗0 = 0 ĵ × k̂

B
c
E
⃗0 = 0 î
B
c
because ĵ × k̂ = î. So the magnetic field plane wave becomes

⃗ E0
B(y, t) = sin(ky − ωt)î
c
So if the wave propagates along the positive-Y direction and the electric field oscillates
along the Z-axis, then the magnetic field oscillates along the X-axis; all directions are mutually
perpendicular.
Let’s summarize our example:

E(y, t) = E0 sin(ky − ωt)k̂

⃗ E0
B(y, t) = sin(ky − ωt)î
c

2.4 Polarization
In what follows, we will only consider and show electric fields. But remember that in all
cases, the magnetic field is also present. Also, in the figures, we will consider vectors only
along the Y-axis. But remember that other vectors are also present.

Department of Physics 30 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
2.4. POLARIZATION

Let us consider an electromagnetic wave propagating in the positive-Y direction. Now,


we know that the electric field oscillates in a direction perpendicular to the direction of
propagation. This direction of oscillation of the electric field is called as the polarization of
the electromagnetic wave. In general, we can write this direction as a combination of two
independent directions: horizontal and vertical. From our perspective, horizontal polarization
is along X-axis, and the vertical polarization is along the Z-axis.

Figure 2.2: Horizontal Polarization

So we can write, mathematically, a pure horizontal oscillation as follows:

E⃗h (y, t) = Eh0 sin(ky − ωt + δh )î

where Eh0 is the amplitude and δh is the phase constant of this horizontally oscillating wave.
Horizontal polarization is as shown in figure 2.2.

Figure 2.3: Vertical Polarization

And we can write, mathematically, a pure vertical oscillation as follows:

E⃗v (y, t) = Ev0 sin(ky − ωt + δv )k̂

where Ev0 is the amplitude and δv is the phase constant of this vertically oscillating wave.
Vertical polarization is as shown in figure 2.3.

Department of Physics 31 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
2.4. POLARIZATION

So any oscillating electric field can be written as a combination of these two polarizations:

E(y, t) = E⃗h (y, t) + E⃗v (y, t)

E(y, t) = Eh0 sin(ky − ωt + δh )î + Ev0 sin(ky − ωt + δv )k̂
Based on the values of the amplitudes and phase constants of these horizontal and vertical
components, we get different types of polarizations.

2.4.1 Linear Polarization

Figure 2.4: Linear Polarization

We get linear polarization when the phase constants of the components are equal, i.e.,
δh = δv and the amplitudes may or may not be equal. Figure 2.4 shows an example of a
linear polarization where δh = 0 = δv .

Figure 2.5: Linear Polarization and components

This is called linear polarization because the path traced by any vector in this field is a
straight line as seen in figure 2.5. In this figure, we see a single vector from the point of view
of a person looking directly at the positive-Y axis and towards the origin. The horizontal
and vertical components as well as the resultant electric field are all shown.

2.4.2 Circular Polarization


We get circular polarization when the difference of phase constants of the components is 90◦ ,
i.e., δh − δv = ±90◦ and the amplitudes are equal, i.e., Eh0 = Ev0 . Figure 2.6 shows an
example of a circular polarization where δh − δv = 90◦ .

Department of Physics 32 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
2.4. POLARIZATION

Figure 2.6: Circular Polarization

Figure 2.7: Circular Polarization and components

This is called circular polarization because the path traced by any vector in this field
is a circle as seen in figure 2.7. The two possible values for the phase difference, 90◦ and
−90◦ , correspond to two different directions for the temporal vector rotation; the former
corresponds to counterclockwise rotation while the latter corresponds to clockwise rotation,
when we choose the convention of looking at the rotation from the direction of propagation
towards the origin.

2.4.3 Elliptical Polarization

We get elliptical polarization when the difference of phase constants of the components is
non-zero, i.e., δh − δv ̸= 0 and the amplitudes may or may not be equal. (Note that circular
polarization is a special case of elliptical polarization.) Figure 2.8 shows an example of an
elliptical polarization where δh − δv = 45◦ .
This is called elliptical polarization because the path traced by any vector in this field is
an ellipse as seen in figure 2.9. Depending on the sign of the phase difference, two different
directions for the temporal vector rotation are possible; positive phase difference corresponds
to anticlockwise rotation while negative phase difference corresponds to clockwise rotation,
when we choose the convention of looking at the rotation from the direction of propagation
towards the origin.

Department of Physics 33 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju
2.5. MODEL AND PREVIOUS YEAR QUESTIONS

Figure 2.8: Elliptical Polarization

Figure 2.9: Elliptical Polarization and components

2.5 Model and Previous Year Questions

Q 1. Derive wave equation in terms of electric field using Maxwell’s equation for free space.
Q 2. Explain the transverse nature of electromagnetic waves.

Department of Physics 34 Physics for EEE Stream Notes


AMC Engineering College BPHYE102/202
Bengaluru-560083 Prof. Nithin Manju

You might also like