Download as pdf or txt
Download as pdf or txt
You are on page 1of 50

EEE 307: ELECROMAGNETIC FIELDS AND WAVES I

VECTOR ANALYSIS
The physic al quantities may be divided into two groups:

1. Scalar: The scalar is a quantity having magnitude but no direction e.g., mass, length and time.
2. Vector: This is a quantity having both magnitude and direction e.g., force, velocity, acceleration,
displacement etc.

A vector is represented by an arrow OP showing the direction, the magnitude of the vector being
indicated by the diagram.

Figure 1:

i.e., a vector is represented by a letter with an arrow over it such as 𝑎̅ 𝑜𝑟 𝑎⃗ and its magnitude is denoted
by |𝑎̅|. Therefore, vector analysis is a mathematical shorthand which facilitates the analysis of electric
and magnetic fields.

Types of vectors
i. Unit vector: is the vector whose magnitude is unity.
ii. Equal vectors: are those vectors which have equal magnitude, same direction (parallel) and
same sense (arrow).
iii. Like vectors: are those vectors which have the same direction (parallel), same sense (arrow)
but the magnitude may be different.
iv. Unlike vectors: are those vectors which have the same direction (parallel), opposite sense
(arrow) but the magnitude may be different.
v. Negative of vector: is a vector whose magnitude is equal to that of the given vector, same
direction (parallel) but opposite sense (arrow).
vi. Zero vector or Null vector: is that vector whose magnitude is zero.

Vector Algebra
1. Vectors may be added or subtracted e.g.,
𝐴̅ = 𝑎𝑥 𝑖 + 𝑎𝑦 𝑗 + 𝑎𝑧 𝑘 and 𝐵̅ = 𝑏𝑥 𝑖 + 𝑏𝑦 𝑗 + 𝑏𝑧 𝑘
𝐴̅ + 𝐵̅ = (𝑎𝑥 𝑖 + 𝑎𝑦 𝑗 + 𝑎𝑧 𝑘) + (𝑏𝑥 𝑖 + 𝑏𝑦 𝑗 + 𝑏𝑧 𝑘)
𝐴̅ + 𝐵̅ = (𝑎𝑥 + 𝑏𝑥 )𝑖 + (𝑎𝑦 + 𝑏𝑦 )𝑗 + (𝑎𝑧 + 𝑏𝑧 )𝑘
and
𝐴̅ − 𝐵̅ = (𝑎𝑥 𝑖 + 𝑎𝑦 𝑗 + 𝑎𝑧 𝑘) − (𝑏𝑥 𝑖 + 𝑏𝑦 𝑗 + 𝑏𝑧 𝑘)
𝐴̅ − 𝐵̅ = (𝑎𝑥 − 𝑏𝑥 )𝑖 + (𝑎𝑦 − 𝑏𝑦 )𝑗 + (𝑎𝑧 − 𝑏𝑧 )𝑘
2. The associative, distributive and commutative laws apply.
𝐴̅ + (𝐵̅ + 𝐶̅ ) = (𝐴̅ + 𝐵̅) + 𝐶̅
𝑘(𝐴̅ + 𝐵̅) = 𝑘𝐴̅ + 𝑘𝐵̅
𝐴̅ + 𝐵̅ = 𝐵̅ + 𝐴̅
3. The dot product (scalar product) of two vectors is by definition
𝐴̅ ∙ 𝐵̅ = 𝐴𝐵 cos 𝜃

Figure 2:
where 𝜃 is a small angle between 𝐴̅ and 𝐵̅.

Example: The dot product obeys the distributive and scalar multiplication laws.

Solution
𝐴̅ ∙ (𝐵̅ + 𝐶̅ ) = 𝐴̅ ∙ 𝐵̅ + 𝐴̅ ∙ 𝐶̅

𝐴̅ ∙ 𝑘𝐵̅ = 𝑘(𝐴̅ ∙ 𝐵̅)

i.e., 𝐴̅ ∙ 𝐵̅ = (𝑎𝑥 𝑖 + 𝑎𝑦 𝑗 + 𝑎𝑧 𝑘) ∙ (𝑏𝑥 𝑖 + 𝑏𝑦 𝑗 + 𝑏𝑧 𝑘)


𝐴̅ ∙ 𝐵̅ = 𝑎𝑥 𝑏𝑥 (𝑖 ∙ 𝑖) + 𝑎𝑥 𝑏𝑦 (𝑖 ∙ 𝑗) + 𝑎𝑥 𝑏𝑧 (𝑖 ∙ 𝑘) + 𝑎𝑦 𝑏𝑥 (𝑗 ∙ 𝑖) + 𝑎𝑦 𝑏𝑦 (𝑗 ∙ 𝑗) + 𝑎𝑦 𝑏𝑧 (𝑗 ∙ 𝑘) + 𝑎𝑧 𝑏𝑥 (𝑘 ∙ 𝑖) + 𝑎𝑧 𝑏𝑦 (𝑘 ∙ 𝑗)
+ 𝑎𝑧 𝑏𝑧 (𝑘 ∙ 𝑘)

However, 𝑖 ∙ 𝑖 = 𝑗 ∙ 𝑗 = 𝑘 ∙ 𝑘 = 1, because the cos 𝜃 in the dot product is unity when the angle
is zero i.e., cos 0 = 1 and 𝑖 ∙ 𝑗 = 𝑖 ∙ 𝑘 = 𝑗 ∙ 𝑖 = 𝑗 ∙ 𝑘 = 𝑘 ∙ 𝑖 = 𝑘 ∙ 𝑗 = 0 is because cos 90∘ = 0.
Thus,

𝐴̅ ∙ 𝐵̅ = 𝑎𝑥 𝑏𝑥 (𝑖 ∙ 𝑖) + 𝑎𝑦 𝑏𝑦 (𝑗 ∙ 𝑗) + 𝑎𝑧 𝑏𝑧 (𝑘 ∙ 𝑘)

𝐴̅ ∙ 𝐵̅ = 𝑎𝑥 𝑏𝑥 + 𝑎𝑦 𝑏𝑦 + 𝑎𝑧 𝑏𝑧

4. The cross product of two vectors is by definition


𝐴̅ × 𝐵̅ = (𝐴𝐵 sin 𝜃)𝑎̅𝑛

Figure 3:

Where 𝜃 is the small angle between 𝐴̅ and 𝐵̅ and 𝑎̅𝑛 is a unit vector normal to the plane
determined by 𝐴̅ and 𝐵̅ when they are drawn from the common point. Because of the
direction requirement, the commutative law does not apply to the cross product; instead

𝐴̅ × 𝐵̅ = −𝐵̅ × 𝐴̅

Figure 4:

Expanding the cross product in component form

𝐴̅ × 𝐵̅ = (𝑎𝑥 𝑖 + 𝑎𝑦 𝑗 + 𝑎𝑧 𝑘) × (𝑏𝑥 𝑖 + 𝑏𝑦 𝑗 + 𝑏𝑧 𝑘)
𝐴̅ × 𝐵̅ = 𝑎𝑥 𝑏𝑥 (𝑖 × 𝑖) + 𝑎𝑥 𝑏𝑦 (𝑖 ∙ 𝑗) + 𝑎𝑥 𝑏𝑧 (𝑖 × 𝑘) + 𝑎𝑦 𝑏𝑥 (𝑗 × 𝑖) + 𝑎𝑦 𝑏𝑦 (𝑗 × 𝑗) + 𝑎𝑦 𝑏𝑧 (𝑗 × 𝑘) + 𝑎𝑧 𝑏𝑥 (𝑘 × 𝑖) + 𝑎𝑧 𝑏𝑦 (𝑘 × 𝑗)
+ 𝑎𝑧 𝑏𝑧 (𝑘 × 𝑘)

𝐴̅ × 𝐵̅ = (𝑎𝑦 𝑏𝑧 − 𝑎𝑧 𝑏𝑦 )𝑖 + (𝑎𝑧 𝑏𝑥 − 𝑎𝑥 𝑏𝑧 )𝑗 + (𝑎𝑥 𝑏𝑦 − 𝑎𝑦 𝑏𝑥 )𝑘


Where 𝑗 × 𝑘 = 𝑖, 𝑘 × 𝑗 = −𝑖, 𝑖 × 𝑘 = 𝑗, 𝑘 × 𝑖 = −𝑗 and 𝑖 × 𝑗 = 𝑘, 𝑗 × 𝑖 = −𝑘. 𝑖 × 𝑖 = 𝑗 × 𝑗 = 𝑘 × 𝑘 = 0.

This can be conveniently expressed as a determinant

𝑖 𝑗 𝑘
𝐴̅ × 𝐵̅ = |𝑎𝑥 𝑎𝑦 𝑎𝑧 |
𝑏𝑥 𝑏𝑦 𝑏𝑧

Example: Given 𝐴̅ = 2𝑖 + 4𝑗 − 3𝑘 and 𝐵̅ = 𝑖 − 𝑗. Find 𝐴̅ ∙ 𝐵̅ and 𝐴̅ × 𝐵̅.

Solution

𝐴̅ ∙ 𝐵̅ = (2𝑖 + 4𝑗 − 3𝑘 ) ∙ (𝑖 − 𝑗)
𝐴̅ ∙ 𝐵̅ = (2)(1)(𝑖 ∙ 𝑖) + (4)(−1)(𝑗 ∙ 𝑗) + (−3)(0)(𝑘 ∙ 𝑘)
𝐴̅ ∙ 𝐵̅ = 2 − 4 + 0 = −2
𝑖 𝑗 𝑘
̅ ̅
𝐴 × 𝐵 = |2 4 −3| = −3𝑖 − 3𝑗 − 6𝑘
1 −1 0

COORDINATE SYSTEMS
A point 𝑃 is described by three coordinates, in cartesian/rectangular (𝑥, 𝑦, 𝑧), in circular cylindrical
(𝑟, ∅, 𝑧) and spherical (𝑟, 𝜃, ∅) as shown below.

Cartesian coordinate Cylindrical coordinate

Spherical coordinate
Figure 5:

The order of specifying the coordinates is important and should be carefully followed. The angle ∅ is the
same angle in both the cylindrical and spherical systems. But in the order of the coordinates, ∅ appears
in the second position in cylindrical (𝑟, ∅, 𝑧), and the third position in spherical (𝑟, 𝜃, ∅). The same
symbol 𝑟 is used in both cylindrical and spherical for two quite different things.
In cylindrical coordinate, 𝑟 measures the distance from the z-axis in a plane normal to the z-axis, while
in the spherical system, 𝑟 measures the distance from the origin 0 to the point 𝑃.
The component forms of a vector in the three systems are
𝐴̅ = 𝑎𝑥 𝑖 + 𝑎𝑦 𝑗 + 𝑎𝑧 𝑘 (cartesian)
𝐴̅ = 𝑎𝑟 𝑖 + 𝑎∅ 𝑗 + 𝑎𝑧 𝑘 (cylindrical)
𝐴̅ = 𝑎𝑟 𝑖 + 𝑎𝜃 𝑗 + 𝑎∅ 𝑘 (spherical)

It should be noted that the components 𝑎𝑥 , 𝑎𝑟 , 𝑎∅ etc. are not generally constants but are functions of
the coordinates in that particular system.

Differential Volume, Surface and Line Elements


There are relatively few problems in electromagnetics that can be solved without some integration
along a curve, over a surface, or through a volume. Hence, the corresponding differential elements must
be clearly understood.
When the coordinates of point 𝑃 are expanded to (𝑥 + 𝑑𝑥, 𝑦 + 𝑑𝑦, 𝑧 + 𝑑𝑧) or (𝑟 + 𝑑𝑟, ∅ + 𝑑∅, 𝑧 + 𝑑𝑧)
or (𝑟 + 𝑑𝑟, 𝜃 + 𝑑𝜃, ∅ + 𝑑∅) a differential volume 𝑑𝑣 is formed.

Cartesian

Cylindrical Spherical
Figure 6:

The variable of the rectangular and cylindrical coordinate systems is easily related to each other with
reference to the figure below.

Figure 7:
hint: 𝑟 = 𝜌
𝑥 = 𝑟 cos ∅ 𝑟 = √𝑥 2 + 𝑦 2
𝑦
𝑦 = 𝑟 sin ∅ ∅ = tan−1 ⁄𝑥
𝑧=𝑧 𝑧=𝑧
Using these equations, scalar functions given in one coordinate system are easily transformed into the
other system. A vector function in one coordinate system, however, requires two steps in order to
transform it to another coordinate system, because a different set of component vectors is generally
required.
We may be given a rectangular vector
𝐴̅ = 𝐴𝑥 𝑎̅𝑥 + 𝐴𝑦 𝑎̅𝑦 + 𝐴𝑧 𝑎̅𝑧

where each component is given as a function of 𝑥, 𝑦, 𝑧 and we need a vector in cylindrical coordinates
𝐴̅ = 𝐴𝑟 𝑎̅𝑟 + 𝐴∅ 𝑎̅∅ + 𝐴𝑧 𝑎̅𝑧

where each component is given as a function of 𝑟, ∅ and 𝑧.


To find the desired component of a vector, we recall from the discussion of the dot product that a
component in a desired direction may be obtained by taking the dot product of the vector and a unit
vector in the desired direction. Hence,
𝐴𝑟 = 𝐴̅ ∙ 𝑎̅𝑟 and 𝐴∅ = 𝐴̅ ∙ 𝑎̅∅

Expanding these dot products, we have


𝐴𝑟 = (𝐴𝑥 𝑎̅𝑥 + 𝐴𝑦 𝑎̅𝑦 + 𝐴𝑧 𝑎̅𝑧 ) ∙ 𝑎̅𝑟 = 𝐴𝑥 𝑎̅𝑥 ∙ 𝑎̅𝑟 + 𝐴𝑦 𝑎̅𝑦 ∙ 𝑎̅𝑟

𝐴∅ = (𝐴𝑥 𝑎̅𝑥 + 𝐴𝑦 𝑎̅𝑦 + 𝐴𝑧 𝑎̅𝑧 ) ∙ 𝑎̅∅ = 𝐴𝑥 𝑎̅𝑥 ∙ 𝑎̅∅ + 𝐴𝑦 𝑎̅𝑦 ∙ 𝑎̅∅

And 𝐴𝑧 = (𝐴𝑥 𝑎̅𝑥 + 𝐴𝑦 𝑎̅𝑦 + 𝐴𝑧 𝑎̅𝑧 ) ∙ 𝑎̅𝑧 = 𝐴𝑧 𝑎̅𝑧 ∙ 𝑎̅𝑧

Since 𝑎̅𝑧 ∙ 𝑎̅𝑟 and 𝑎̅𝑧 ∙ 𝑎̅∅ are zero.


In order to complete the transformation of the components, it is necessary to know the dot products
𝑎̅𝑥 ∙ 𝑎̅𝑟 , 𝑎̅𝑦 ∙ 𝑎̅𝑟 , 𝑎̅𝑥 ∙ 𝑎̅∅ and 𝑎̅𝑦 ∙ 𝑎̅∅ . Applying the definition of the dot product, we see that since we are
concerned with unit vectors, the result is merely the cosine of the angle between the two-point unit
vectors.

GRADIENT, DIVERGENCE AND CURL OPERATION


In Calculus, a gradient is a term used for the differential operator, which is applied to the three-
dimensional vector-valued function to generate a vector. The symbol used to represent the gradient is
∇ (nabla). For example, if ′𝑓′ is a function, then the gradient of a function is represented by ′∇𝑓′. In this
article, let us discuss the definition gradient of a function, directional derivative, properties and solved
examples in detail.
The gradient of a function is defined to be a vector field. Generally, the gradient of a function can be
found by applying the vector operator to the scalar function ∇𝑓(𝑥, 𝑦, 𝑧). This kind of vector field is
known as the gradient vector field.

Gradient of Function in Two and Three Dimensions


If the function is 𝑓(𝑥, 𝑦), then the gradient of a function is 2-dimensional and is given by:
𝛿 𝛿
𝑔𝑟𝑎𝑑 𝑓(𝑥, 𝑦) = ∇𝑓(𝑥, 𝑦) = ( 𝑖 + 𝑗) 𝑓(𝑥, 𝑦)
𝛿𝑥 𝛿𝑦

If the function is 𝑓(𝑥, 𝑦, 𝑧), then the gradient of a function is 3-dimensional and is given by:
𝛿 𝛿 𝛿
𝑔𝑟𝑎𝑑 𝑓(𝑥, 𝑦, 𝑧) = ∇𝑓(𝑥, 𝑦, 𝑧) = ( 𝑖 + 𝑗 + 𝑘) 𝑓(𝑥, 𝑦, 𝑧)
𝛿𝑥 𝛿𝑦 𝛿𝑧
Directional Derivative
The component of the gradient of the function (∇𝑓) in any direction is defined as the rate of change of
the function in that direction. For example, the component in ′𝑖′ direction is the partial derivative of the
function with respect to 𝑥. In other words, we can say that it is the rate of change of function in the 𝑥-
direction, by keeping 𝑦 and 𝑧 as constant.

Properties of Gradient
The following are the important properties of the gradient of a function:
1. The gradient should take a scalar function (i.e., 𝑓(𝑥, 𝑦) and produces the vector function (∇𝑓).
2. The vector ∇𝑓(𝑥, 𝑦) should lie in the plane.

Example: Find the gradient of function 𝑓(𝑥, 𝑦) = 𝑥 + 3𝑦 2

Solution
Given function: 𝑓(𝑥, 𝑦) = 𝑥 + 3𝑦 2
𝛿 𝛿
We know that, ∇𝑓(𝑥, 𝑦) = (𝛿𝑥 𝑖 + 𝛿𝑦 𝑗) 𝑓(𝑥, 𝑦)

𝛿 𝛿
∇𝑓(𝑥, 𝑦) = ( 𝑖 + 𝑗) (𝑥 + 3𝑦 2 )
𝛿𝑥 𝛿𝑦

𝛿(𝑥 + 3𝑦 2 ) 𝛿(𝑥 + 3𝑦 2 )
∇𝑓(𝑥, 𝑦) = 𝑖+ 𝑗 = (1 + 0)𝑖 + (0 + 6𝑦)𝑗
𝛿𝑥 𝛿𝑦

∇𝑓(𝑥, 𝑦) = 𝑖 + 6𝑦𝑗

Example: Find the gradient of the function, ∇ 𝑓(𝑥, 𝑦, 𝑧) = sin(𝑥) 𝑒 𝑦 ln(𝑧)

Solution
Given function: ∇𝑓(𝑥, 𝑦, 𝑧) = sin(𝑥) 𝑒 𝑦 ln(𝑧)
𝛿 𝛿 𝛿
As we know, ∇𝑓(𝑥, 𝑦, 𝑧) = (𝛿𝑥 𝑖 + 𝛿𝑦 𝑗 + 𝛿𝑧 𝑘) 𝑓(𝑥, 𝑦, 𝑧)

𝛿𝑓 𝛿 sin(𝑥)
𝛿𝑥
= 𝑒 𝑦 ln(𝑧) 𝛿𝑥
= cos(𝑥) 𝑒 𝑦 ln(𝑧)
𝛿𝑓 𝛿 𝑒𝑦
= sin(𝑥) ln(𝑧) = sin(𝑥) 𝑒 𝑦 ln(𝑧)
𝛿𝑥 𝛿𝑥
𝛿𝑓 𝛿 ln(𝑧) 1
𝛿𝑥
= sin(𝑥) 𝑒 𝑦 𝛿𝑥 = sin(𝑥) 𝑒 𝑦 𝑧

1
∇𝑓(𝑥, 𝑦, 𝑧) = cos(𝑥) 𝑒 𝑦 ln(𝑧) 𝑖 + sin(𝑥) 𝑒 𝑦 ln(𝑧) 𝑗 + sin(𝑥) 𝑒 𝑦 𝑘
𝑧

Divergence of Vector Field


The divergence of a vector field is a scalar field. The divergence is generally denoted by ′𝑑𝑖𝑣′. The
divergence of a vector field can be calculated by taking the scalar product of the vector operator applied
to the vector field i.e., ∇ ∙ 𝐹(𝑥, 𝑦)
If 𝐹(𝑥, 𝑦) is a vector field in the two dimensions, then its divergence is given by:
𝛿 𝛿 𝛿𝐹1 (𝑥,𝑦) 𝛿𝐹 (𝑥,𝑦)
∇ ∙ 𝐹(𝑥, 𝑦) = ( 𝑖 + 𝑗) ∙ 𝐹(𝑥, 𝑦) = + 2
𝛿𝑥 𝛿𝑦 𝛿𝑥 𝛿𝑦
The divergence of a vector field can be extended to three dimensions and it is given as follows:
i.e., 𝐹(𝑥, 𝑦, 𝑧) = 𝐹1 𝑖 + 𝐹2 𝑗 + 𝐹3 𝑘
𝛿 𝛿 𝛿 𝛿𝐹1 𝛿𝐹2 𝛿𝐹3
∇ ∙ 𝐹(𝑥, 𝑦, 𝑧) = ( 𝑖 + 𝑗 + 𝑘) ∙ 𝐹(𝑥, 𝑦, 𝑧) = + +
𝛿𝑥 𝛿𝑦 𝛿𝑧 𝛿𝑥 𝛿𝑦 𝛿𝑧

Example: Determine the divergence of a vector field in two dimensions: 𝐹(𝑥, 𝑦) = 6𝑥 2 𝑖 + 4𝑦𝑗

Solution
Given: 𝐹(𝑥, 𝑦) = 6𝑥 2 𝑖 + 4𝑦𝑗
𝛿𝐹1 (𝑥,𝑦) 𝛿𝐹 (𝑥,𝑦)
We know that, ∇ ∙ 𝐹(𝑥, 𝑦) = + 2
𝛿𝑥 𝛿𝑦

𝛿 (6𝑥 2 ) 𝛿(4𝑦)
∇ ∙ 𝐹(𝑥, 𝑦) = + = 12𝑥 + 4
𝛿𝑥 𝛿𝑦

Example: Find the divergence of a vector field in three dimensions: 𝐹(𝑥, 𝑦, 𝑧) = 𝑥 2 𝑖 + 2𝑧𝑗 − 𝑦𝑘

Solution
Given: 𝐹(𝑥, 𝑦, 𝑧) = 𝑥 2 𝑖 + 2𝑧𝑗 − 𝑦𝑘
𝛿𝐹1 𝛿𝐹2 𝛿𝐹3
Recall that, ∇ ∙ 𝐹(𝑥, 𝑦, 𝑧) = + +
𝛿𝑥 𝛿𝑦 𝛿𝑧

𝛿(𝑥 2 ) 𝛿(2𝑧) 𝛿(−𝑦)


∇ ∙ 𝐹(𝑥, 𝑦, 𝑧) = 𝛿𝑥
+ 𝛿𝑦
+ 𝛿𝑦
= 2𝑥 + 0 + 0 = 2𝑥

Curl of a Vector Field


The curl of a vector field is again a vector field. The curl of a vector field is obtained by taking the vector
product of the vector operator applied to the vector field 𝐹(𝑥, 𝑦, 𝑧).
i.e., 𝑐𝑢𝑟𝑙 𝐹(𝑥, 𝑦, 𝑧) = ∇ × 𝐹(𝑥, 𝑦, 𝑧)
It can also be written as:
𝛿𝐹3 𝛿𝐹2 𝛿𝐹3 𝛿𝐹1 𝛿𝐹2 𝛿𝐹1
∇ × 𝐹(𝑥, 𝑦, 𝑧) = ( − )𝑖 − ( − )𝑗 + ( + )𝑘
𝛿𝑦 𝛿𝑧 𝛿𝑥 𝛿𝑧 𝛿𝑥 𝛿𝑦

𝑖 𝑗 𝑘
𝛿 𝛿 𝛿
∇ × 𝐹(𝑥, 𝑦, 𝑧) = |𝛿𝑥 𝛿𝑦 𝛿𝑧
|
𝐹1 𝐹2 𝐹3

Example: Find the curl of the vector field 𝐹(𝑥, 𝑦, 𝑧) = 𝑦 3 𝑖 + 𝑥𝑦𝑗 − 𝑧𝑘

Solution
Given: 𝐹(𝑥, 𝑦, 𝑧) = 𝑦 3 𝑖 + 𝑥𝑦𝑗 − 𝑧𝑘
Here, 𝐹1 = 𝑦 3 , 𝐹2 = 𝑥𝑦, 𝐹3 = −𝑧

𝛿𝐹 𝛿𝐹2 𝛿𝐹 𝛿𝐹1 𝛿𝐹 𝛿𝐹1


Recall that, ∇ × 𝐹(𝑥, 𝑦, 𝑧) = ( 𝛿𝑦3 − 𝛿𝑧
)𝑖 − ( 𝛿𝑥3 − 𝛿𝑧
)𝑗 + ( 𝛿𝑥2 + 𝛿𝑦
)𝑘

𝛿(−𝑧) 𝛿(𝑥𝑦) 𝛿(−𝑧) 𝛿(𝑦 3 ) 𝛿(𝑥𝑦) 𝛿(𝑦 3 )


∇ × 𝐹(𝑥, 𝑦, 𝑧) = ( − )𝑖 − ( − )𝑗 + ( − )𝑘
𝛿𝑦 𝛿𝑧 𝛿𝑥 𝛿𝑧 𝛿𝑥 𝛿𝑦

∇ × 𝐹(𝑥, 𝑦, 𝑧) = 0𝑖 − 0𝑗 + (𝑦 − 3𝑦 2 )𝑘

∇ × 𝐹(𝑥, 𝑦, 𝑧) = (𝑦 − 3𝑦 2 )𝑘
Therefore, the curl of the vector is in the direction of ′𝑘′.

Exercise
1. Solve the following problems.
a. Compute the divergence of the vector field 𝐹(𝑥, 𝑦, 𝑧) = 𝑦 3 𝑖 + 𝑥𝑦𝑧 𝑗 − 𝑥𝑧𝑒 𝑦𝑧 𝑘.
b. Calculate the divergence of the vector field 𝐹(𝑥, 𝑦, 𝑧) = sin(𝑥) 𝑖 − 𝑒 𝑦 𝑗 + ln(𝑧) 𝑘.
c. Find the curl of the vector field 𝐹(𝑥, 𝑦, 𝑧) = 𝑦 3 𝑖 + 𝑥𝑦 𝑗 − 𝑥𝑧 𝑘.
2. Prove the following:
a. ∇ ∙ (∇ × 𝐹) = 0
b. ∇ × (∇𝑓) = 0
GAUSS’S, STOKE’S AND GREEN’S THEOREMS
This section finally begins to deliver on why we introduced 𝑑𝑖𝑣 𝑔𝑟𝑎𝑑 and 𝑐𝑢𝑟𝑙. Two theorems, both of
them over two hundred years old, are explained:

Gauss’s Theorem enables an integral taken over a volume to be replaced by one taken over the surface
bounding that volume, and vice versa. Why would we want to do that? Computational efficiency and/or
numerical accuracy!

Stoke’s Law enables an integral taken around a closed curve to be replaced by one taken over any
surface bounded by that curve.

Gauss’s Theorem
Suppose that 𝑎(𝑟) is a vector field and we want to compute the total flux of the field across the surface
𝑆 that bounds a volume 𝑉 . That is, we are interested in calculating:

∫ 𝑎 ∙ 𝑑𝑆
𝑆
where recall that 𝑑𝑆 is normal to the locally planar surface element and must everywhere point out of
the volume as shown in Figure 8.

Figure 8: The surface element 𝑑𝑆 must stick out of the surface.

Gauss’s theorem tells us that we can do this by considering the total flux generated inside the volume
𝑉:

∫ 𝑎 ∙ 𝑑𝑆 = ∫ 𝑑𝑖𝑣 𝑎 ∙ 𝑑𝑉
𝑆 𝑉
obtained by integrating the divergence over the entire volume.

Proof
A non-rigorous proof can be realized by recalling that we defined 𝑑𝑖𝑣 by considering the efflux 𝑑𝐸 from
the surfaces of an infinitesimal volume element
𝑑𝐸 = 𝑎 ∙ 𝑑𝑆
and defining it as
𝑑𝑖𝑣 𝑎 ∙ 𝑑𝑉 = 𝑑𝐸 = 𝑎 ∙ 𝑑𝑆

If we sum over the volume elements, this results in a sum over the surface elements. But if two
elemental surface touch, their 𝑑𝑆 vectors are in opposing direction and cancel as shown in Figure. Thus,
the sum over surface elements gives the overall bounding surface.
This is a typical example, in which the surface integral is rather tedious, whereas the volume integral is
straightforward.
Figure 9: When two elements touch, the 𝑑𝑆 vectors at the common surface cancel out. One can imagine
building the entire volume up from the infinitesimal units.

Example: Derive ∫𝑆 𝑎 ∙ 𝑑𝑆 where 𝑎 = 𝑧 3 𝑘̂ and 𝑆 is the surface of a sphere of radius 𝑅 centred at the
origin:
i. directly;
ii. by applying Gauss’ Theorem

Figure 10:

Solution
i. On the surface of the sphere, 𝑎 = 𝑧 3 𝑘̂ = 𝑅 3 sin3 𝜃 𝑘̂ and 𝑑𝑆 = 𝑅 2 sin 𝜃 𝑑𝜃 𝑑𝜙 𝑟̂ .
Everywhere 𝑟̂ ∙ 𝑘̂ = cos 𝜃.
2𝜋 𝜋
∫ 𝑎 ∙ 𝑑𝑆 = ∫ ∫ 𝑅 3 sin3 𝜃 ∙ 𝑅 2 sin 𝜃 𝑑𝜃 𝑑𝜙 𝑟̂ ∙ 𝑘̂
𝑆 𝜙=0 𝜃=0

2𝜋 𝜋
= ∫𝜙=0 ∫𝜃=0 𝑅 3 sin3 𝜃 ∙ 𝑅 2 sin 𝜃 𝑑𝜃 𝑑𝜙 ∙ cos 𝜃

𝜋
= 2𝜋𝑅 5 ∫𝜃=0 cos 4 𝜃 sin 𝜃 𝑑𝜃

2𝜋𝑅5 4𝜋𝑅5
= [− cos4 𝜃]𝜋0 =
5 5

ii. To apply Gauss’ Theorem, we need to figure out 𝑑𝑖𝑣 𝑎 and decide how to compute the
volume integral. The first is easy:
𝑑𝑖𝑣 𝑎 = 3𝑧 2
For the second, because 𝑑𝑖𝑣 𝑎 involves just 𝑧, we can divide the sphere into discs of
constant 𝑧 and thickness 𝑑𝑧, as shown in Figure 10. Then
𝑑𝑉 = 𝜋(𝑅 2 − 𝑧 2 )𝑑𝑧

𝑅
∫ 𝑑𝑖𝑣 𝑎 𝑑𝑉 = 3𝜋 ∫ 𝑧 2 (𝑅2 − 𝑧 2 )𝑑𝑧
𝑉 −𝑅

𝑅
𝑧 3 𝑅2 𝑧 5 4𝜋𝑅 5
= 3𝜋 [ − ] =
3 5 −𝑅 5
Extension of Gauss’s Theorem
Suppose the vector field 𝑎(𝑟) is of the form 𝑎 = 𝑈(𝑟)𝑐, where U(r) as scalar field and 𝑐 is a constant
vector.
𝑑𝑖𝑣 𝑎 = 𝑔𝑟𝑎𝑑𝑈 ∙ 𝑐 + 𝑈 𝑑𝑖𝑣 𝑐

𝑑𝑖𝑣 𝑎 = 𝑔𝑟𝑎𝑑𝑈 ∙ 𝑐
Since 𝑑𝑖𝑣 𝑐 = 0 because 𝑐 is a constant. Then Gauss’ Theorem becomes

∫ 𝑈𝑐 ∙ 𝑑𝑆 = ∫ 𝑔𝑟𝑎𝑑 𝑈 ∙ 𝑐𝑑𝑉
𝑆 𝑉
or, alternatively, taking the constant 𝑐 out of the integrals

𝑐 ∙ (∫ 𝑈𝑑𝑆) = 𝑐 ∙ (∫ 𝑔𝑟𝑎𝑑 𝑈𝑑𝑉 )


𝑆 𝑉
This is still a scalar equation but we now note that the vector 𝑐 is arbitrary so that the result must be
true for any vector 𝑐. This can be true only if the vector equation

∫ 𝑈𝑑𝑆 = ∫ 𝑔𝑟𝑎𝑑 𝑈𝑑𝑉


𝑆 𝑉
is satisfied.
If you think this is fishy, just write 𝑐 = 𝑖̂, then 𝑐 = 𝑗̂ and 𝑐 = 𝑘̂ in turn, and you must obtain the three
components of ∫𝑆 𝑈𝑑𝑆 in turn.
Further “extensions” can be obtained of course. For example, one might be able to write the vector field
of interest as
𝑎(𝑟) = 𝑏(𝑟) × 𝑐
Where 𝑐 is a constant vector.

Example: 𝑈 = 𝑥 2 + 𝑦 2 + 𝑧 2 is a scalar field, and volume 𝑉 is a cylinder 𝑥 2 + 𝑦 2 ≤ 𝑎2 , 0 ≤ 𝑧 ≤ ℎ.


Compute the surface integral using Gauss’s theorem.

Solution

∫ 𝑈𝑑𝑆 = ∫ 𝑔𝑟𝑎𝑑 𝑈𝑑𝑉


𝑆 𝑉
𝑈 = 𝑥 2 + 𝑦 2 + 𝑧 2 = 𝑟 2 , 𝑔𝑟𝑎𝑑 𝑈 = 2𝑟, 𝑑𝑉 = 𝑟𝑑𝑟 𝑑𝜙 𝑑𝑧
𝑟 = 𝑥𝑖̂ + 𝑦𝑗̂ + 𝑧𝑘̂

= 2 ∫ (𝑥𝑖̂ + 𝑦𝑗̂ + 𝑧𝑘̂ ) 𝑟𝑑𝑟 𝑑𝜙 𝑑𝑧


𝑉
The integrations over 𝑥 and 𝑦 are zero by symmetry, so that the only remaining part is
ℎ 𝑎 2𝜋

= 2 ∫ 𝑧𝑑𝑧 + ∫ 𝑟𝑑𝑟 + ∫ 𝑑𝜙𝑘̂ = 𝜋𝑎2 ℎ2 𝑘̂


𝑧=0 𝑟=0 𝜙=0

Stoke’s Theorem
Stoke’s Theorem relates a line integral around a closed path to a surface integral over what is called a
capping surface of the path.
Stokes’ Theorem states: ∮𝐶 𝑎 ∙ 𝑑𝑙 = ∫𝑆 𝑐𝑢𝑟𝑙 𝑎 ∙ 𝑑𝑆. where 𝑆 is any surface capping the curve 𝐶.

Why have we used 𝑑𝑙 rather than 𝑑𝑟, where 𝑟 is the position vector?
There is no good reason for this, as 𝑑𝑙 = 𝑑𝑟. It just seems to be common usage in line integrals!

Proof
You will recall that we defined 𝑐𝑢𝑟𝑙 as the circulation per unit area, and showed that 𝑋 around elemental
loop
∑ 𝑎 ∙ 𝑑𝑙 = 𝑑𝐶 = (∇ × 𝑎) ∙ 𝑑𝑆
𝑎𝑟𝑜𝑢𝑛𝑑 𝑡ℎ𝑒 𝑒𝑙𝑒𝑚𝑒𝑛𝑡𝑎𝑙 𝑙𝑜𝑜𝑝
Now, if we add these little loops together, the internal line sections cancel out because the 𝑑𝑙’s are in
opposite direction but the field 𝑎 is not. This gives the larger surface and the larger bounding contour
as shown in Figure 11.

Figure 11: An example of an elementary loop, and how they combine together.

For a given contour, the capping surface can be ANY surface bound by the contour. The only requirement
is that the surface element vectors point in the “general direction” of a right-handed screw with respect
to the sense of the contour integral. See Figure 12.

Figure 12: For a given contour, the bounding surface can be any shape. 𝑑𝑆’s must have a positive
component in the sense of a r-h screw w.r.t the contour sense.

Example: Let 𝐹 = 𝑥𝑒 𝑧 𝑖̂ + (𝑥 + 𝑥𝑧)𝑗̂ + 3𝑒 𝑧 𝑘̂ and let 𝑆 be the top half of the sphere 𝑥 2 + 𝑦 2 + 𝑧 2 = 1.
Find ∫𝑆 (𝛻 × 𝐹) ∙ 𝑛 𝑑𝑆 where 𝑛 is outward normal.

Solution

∫ (𝛻 × 𝐹) ∙ 𝑛 𝑑𝑆 = ∫ (𝛻 × 𝐹) ∙ 𝑘̂ 𝑑𝑆
𝑆 𝑆∗
(On 𝑆 ∗ note that 𝑘, not −𝑘, is the correct normal to use.)
𝑖̂ 𝑗̂ 𝑘̂
𝛿 𝛿 𝛿
𝛻×𝐹 =| |
𝛿𝑥 𝛿𝑦 𝛿𝑧
𝑥𝑒 𝑧 𝑥 + 𝑥𝑧 3𝑒 𝑧

𝛿(3𝑒 𝑧 ) 𝛿(𝑥 + 𝑥𝑧) 𝛿(3𝑒 𝑧 ) 𝛿(𝑥𝑒 𝑧 ) 𝛿(𝑥 + 𝑥𝑧) 𝛿(𝑥𝑒 𝑧 )


𝛻×𝐹 =( − ) 𝑖̂ − ( − ) 𝑗̂ + ( − ) 𝑘̂
𝛿𝑦 𝛿𝑧 𝛿𝑥 𝛿𝑧 𝛿𝑥 𝛿𝑦

𝛻 × 𝐹 = −𝑥 𝑖̂ + 𝑥𝑒 𝑧 𝑗̂ + (1 + 𝑧) 𝑘̂

(𝛻 × 𝐹) ∙ 𝑘̂ = (−𝑥 𝑖̂ + 𝑥𝑒 𝑧 𝑗̂ + (1 + 𝑧) 𝑘̂ ) ∙ 𝑘̂ = 1 + 𝑧
On 𝑆 ∗, 𝑧 = 0, so

∫ (𝛻 × 𝐹) ∙ 𝑘 𝑑𝑆 = ∫ (1 + 0) 𝑑𝑆
𝑆∗ 𝑆∗
2𝜋
∫ 𝑑𝑆 = 2𝜋
0
Thus, the original integral over 𝑆 is also 2𝜋.
Example: Vector field 𝑎 = 𝑥 3 𝑗̂ − 𝑦 3 𝑖̂ and 𝐶 is the circle of radius 𝑅 centred on the origin. Derive
∮𝐶 𝑎 ∙ 𝑑𝑙 using the Stoke’s theorem, where the surface is the planar surface bounded by the contour.

Solution

∮ 𝑎 ∙ 𝑑𝑙 = ∫ (𝛻 × 𝑎) ∙ 𝑑𝑆
𝑆
𝐶
𝑖̂ 𝑗̂ 𝑘̂
𝛿 𝛿 𝛿
𝛻 × 𝑎 = || |
𝛿𝑥 𝛿𝑦 𝛿𝑧|
−𝑦 3 𝑥3 0

𝛿(0) 𝛿(𝑥 3 ) 𝛿(0) 𝛿(−𝑦 3 ) 𝛿(𝑥 3 ) 𝛿(−𝑦 3 )


𝛻×𝑎 =( − ) 𝑖̂ − ( − ) 𝑗̂ + ( − ) 𝑘̂
𝛿𝑦 𝛿𝑧 𝛿𝑥 𝛿𝑧 𝛿𝑥 𝛿𝑦

𝛻 × 𝑎 = (3𝑥 2 + 3𝑦 2 )𝑘̂ = 3(𝑥 2 + 𝑦 2 )𝑘̂

𝛻 × 𝑎 = 3𝑟 2 𝑘̂
2 2 2
Where 𝑟 = 𝑥 + 𝑦 , We choose area elements to be circular strips of radius 𝑟 with thickness 𝑑𝑟, then,
𝑑𝑆 = 2𝜋𝑟 𝑑𝑟 𝑘̂

∫ (𝛻 × 𝑎) ∙ 𝑑𝑆 = ∫ 3𝑟 2 𝑘̂ ∙ 2𝜋𝑟 𝑑𝑟 𝑘̂
𝑆 𝑆

𝑅 𝑅
3
𝑟4 3𝜋𝑅 4
= 6𝜋 ∫ 𝑟 𝑑𝑟 = 6𝜋 [ ] =
0 4 0 2

An Extension to Stoke’s Theorem


Just as we considered one extension to Gauss’s theorem (not really an extension, more of a re-
expression), so we will try something similar with Stoke’s theorem.

Let 𝑎(𝑟) = 𝑈(𝑟)𝑐 where 𝑐 is a constant vector. Then,


𝑐𝑢𝑟𝑙 𝑎 = 𝑈 𝑐𝑢𝑟𝑙 𝑐 + 𝑔𝑟𝑎𝑑𝑈 × 𝑐

∇ × 𝑎 = 𝑈 (∇ × 𝑐) + ∇𝑈 × 𝑐
Where ∇ × 𝑐 = 0, therefore, Stoke’s theorem becomes

∮ 𝑈(𝑐 ∙ 𝑑𝑙) = ∫ (∇𝑈 × 𝑐 ∙ 𝑑𝑆) = ∫ 𝑐 ∙ (𝑑𝑆 × ∇𝑈)


𝑆 𝑆
𝐶
or, rearranging the triple scalar products and taking the constant 𝑐 out of the integrals gives

𝑐 ∙ ∮ 𝑈𝑑𝑙 = −𝑐 ∙ ∫ (∇𝑈 × 𝑑𝑆)


𝑆
𝐶
But 𝑐 is arbitrary and so,

∮ 𝑈𝑑𝑙 = − ∫ (∇𝑈 × 𝑑𝑆)


𝑆
𝐶

Example: Derive ∮𝐶 𝑈𝑑𝑟 using Stoke’s, where 𝑈 = 𝑥 2 + 𝑦 2 + 𝑧 2 and the line integral is taken around
𝐶 the circle (𝑥 − 𝑎)2 + 𝑦 2 = 𝑎2 and 𝑧 = 0. Given that 𝑥 = 𝑎 + 𝜌 cos 𝛼 and 𝑦 = 𝜌 sin 𝛼.

Solution
For a planar surface covering the disc, the surface element can be written using the new parametrization
as
𝑑𝑆 = 𝜌 𝑑𝜌 𝑑𝛼 𝑘̂

Remember that 𝑈 = 𝑥 2 + 𝑦 2 + 𝑧 2 = 𝑟 2 and 𝑧 = 0 in the plane.

𝛿 𝛿 𝛿
∇𝑈 = ( 𝑖̂ + 𝑗̂ + 𝑘̂ ) (𝑥 2 + 𝑦 2 + 𝑧 2 )
𝛿𝑥 𝛿𝑦 𝛿𝑧

∇𝑈 = 2(𝑥 𝑖̂ + 𝑦 𝑗̂ + 𝑧 𝑘̂ ) = 2(𝑎 + 𝜌 cos 𝛼)𝑖̂ + 2(𝜌 sin 𝛼)𝑗̂

Be careful to note that 𝑥, 𝑦 are specified for any point on the disc, not on its circular boundary! So

𝑖̂ 𝑗̂ 𝑘̂
𝑑𝑆 × ∇𝑈 = 2𝜌 𝑑𝜌 𝑑𝛼 | 0 0 1|
𝑎 + 𝜌 cos 𝛼 𝜌 sin 𝛼 0

𝑑𝑆 × ∇𝑈 = 2𝜌(−𝜌 sin 𝛼 𝑖̂ + (𝑎 + 𝜌 cos 𝛼)𝑗̂) 𝑑𝜌 𝑑𝛼

𝑎 2𝜋

∫ 𝑑𝑆 × ∇𝑈 = ∫ ∫ 2𝜌(−𝜌 sin 𝛼 𝑖̂ + (𝑎 + 𝜌 cos 𝛼)𝑗̂) 𝑑𝜌 𝑑𝛼 = 2𝜋𝑎3 𝑗̂


𝑆
𝜌=0 𝛼=0

2𝜋 2𝜋
Both ∫0 sin 𝛼 𝑑𝛼 = 0 and ∫0 cos 𝛼 𝑑𝛼 = 0, so we have

𝑎 2𝜋

∫ 𝑑𝑆 × ∇𝑈 = ∫ ∫ 2𝜌𝑎 𝑗̂ 𝑑𝜌 𝑑𝛼 = 2𝜋𝑎3 𝑗̂
𝑆
𝜌=0 𝛼=0

Green's Theorem
In this section we discuss a theorem that relates an integral of a vector field over a closed curve 𝐶 in a
plane to an integral of a related scalar function over the region 𝑅 whose boundary is 𝐶.

Let 𝐶 be a simple, closed counter clockwise curve in the 𝑥𝑦-plane, bounding a region 𝑅. Let 𝑃 and 𝑄 be
scalar functions defined at least on an open set containing 𝑅. Assume 𝑃 and 𝑄 have continuous first
partial derivatives. Then
𝜕𝑄 𝜕𝑃
∮ (𝑃 𝑑𝑥 + 𝑄 𝑑𝑦) = − ∫ ( − ) 𝑑𝐴
𝑅 𝜕𝑥 𝜕𝑦
𝐶

Figure 13
It could be seen from the figure that the curve is closed when it starts and ends at the same point. It's
simple when it does not intersect itself (except at its start and end). These restrictions on 𝐶 ensure that
it is the boundary of a region 𝑅 in the 𝑥𝑦 -plane.
Since 𝑃 and 𝑄 are independent of each other, Green's Theorem really consists of two theorems:
𝜕𝑃
∮ 𝑃 𝑑𝑥 = ∫ 𝑑𝐴
𝜕𝑦
𝐶 𝑅
and
𝜕𝑄
∮ 𝑄 𝑑𝑦 = ∫ 𝑑𝐴
𝜕𝑥
𝐶 𝑅
Green's Theorem and Circulation
What does Green's Theorem say about a vector field 𝐹 = 𝑃𝑖̂ + 𝑄𝑗̂ First of all, ∮𝐶 (𝑃 𝑑𝑥 + 𝑄 𝑑𝑦) now
becomes simply ∮𝐶 𝐹 𝑑𝑟.
The right-hand side of Green's Theorem looks a bit like the curl of a vector field in the plane. To be
specific, we compute the curl of 𝐹:
𝑖̂ 𝑗̂ 𝑘̂
𝛻 × 𝐹 = | 𝜕𝑥 𝜕𝑦 𝜕𝑧|
𝑃(𝑥, 𝑦) 𝑄(𝑥, 𝑦) 0

𝜕𝑄 𝜕𝑃
𝛻 × 𝐹 = 0𝑖̂ − 0𝑗̂ + ( − ) 𝑘̂
𝜕𝑥 𝜕𝑦

𝜕𝑄 𝜕𝑃
𝛻×𝐹 =( − ) 𝑘̂
𝜕𝑥 𝜕𝑦
To obtain the (scalar) integrand on the right-hand side, we dot with 𝑘̂ i.e.,

𝜕𝑄 𝜕𝑃 𝜕𝑄 𝜕𝑃
(( − ) 𝑘̂ ) ∙ 𝑘̂ = −
𝜕𝑥 𝜕𝑦 𝜕𝑥 𝜕𝑦

Example: Solve ∮𝐶 𝑦 3 𝑑𝑥 − 𝑥 3 𝑑𝑦, where 𝐶 is a circle of radius 2 cantered at the origin.

Solution
Given ∮𝐶 𝑦 3 𝑑𝑥 − 𝑥 3 𝑑𝑦
Let’s identify 𝑃 amd 𝑄 from the integral
𝑃 = 𝑦 3 and 𝑄 = −𝑥 3
Recall that
𝜕𝑄 𝜕𝑃
∮ (𝑃 𝑑𝑥 + 𝑄 𝑑𝑦) = − ∫ ( − ) 𝑑𝐴
𝑅 𝜕𝑥 𝜕𝑦
𝐶

𝜕(−𝑥 3 ) 𝜕(𝑦 3 )
∮ (𝑦 3 𝑑𝑥 − 𝑥 3 𝑑𝑦) = − ∫ ( − ) 𝑑𝐴
𝑅 𝜕𝑥 𝜕𝑦
𝐶

= − ∫ (−3𝑥 2 − 3𝑦 2 ) 𝑑𝐴
𝑅
𝑑𝐴 = 𝑟 𝑑𝑟 𝑑𝜃 since the problem under study is circular

= −3 ∫ (−𝑥 2 − 𝑦 2 ) 𝑑𝐴
𝑅
Let 𝑥 2 + 𝑦 2 = 𝑟 2
2𝜋 2

= −3 ∫ ∫ 𝑟 2 ∙ 𝑟𝑑𝑟 𝑑𝜃
0 0

2𝜋 2 2𝜋 2
3
𝑟4
= −3 ∫ 𝑑𝜃 ∫ 𝑟 𝑑𝑟 = −3 ∫ | | 𝑑𝜃
4 0
0 0 0

2𝜋

= −3 ∫ 4 𝑑𝜃 = −3|4𝜃|2𝜋
0 = −3(8𝜋) = −24𝜋
0
COULOMB’S LAW
Colonel Charles Coulomb (an officer in the French army engineer) performed an elaborate series of
experiments using a delicate torsion balance, invented by himself, to determine quantitatively the force
exerted between two objects each having a static charge of electricity. His published result is now known
to many high school students and bears a great similarity to newton’s gravitational law (discovered
about a hundred years earlier).

Coulomb force law states that the magnitude of electrostatic force between two-point charge separated
in a vacuum or free space is proportional to the product of the magnitude of the charges in each and
inversely proportional to the square of the distance between them.

𝑄1 𝑄2
𝐹=𝑘 𝑅2
1

Where 𝑄1 and 𝑄2 are the positive or negative quantities of charge, 𝑅 is the separation and 𝑘 is the
proportionality constant.
If the International System (SI) of unit is used, 𝑄 is measured in Coulombs (𝐶), 𝑅 is measured in meters
1
(𝑚) and 𝐹 is measured in Newton (𝑁), the constant of proportionality 𝑘 = . The constant 𝜀𝑜 is
4𝜋𝜀𝑜
called the permittivity of free space and has the magnitude measured in Farads per meter (𝐹/𝑚).
1
𝜀𝑜 = 8.854 × 10−12 = × 10−9 𝐹/𝑚 2
36𝜋

Therefore, Coulomb’s law is now


𝑄 𝑄2
𝐹 = 4𝜋𝜀1 2 3
𝑜𝑅

In order to write the vector form of equation (3), we need the additional fact (furnished also by Colonel
Coulomb) that the force acts along the line joining the two charges and is repulsive if the charges are
alike in sign and attractive if they are of opposite sign. Let the vector 𝑟̅1 locate 𝑄1 while 𝑟̅2 locate 𝑄2 .
Then the vector 𝑅̅12 = 𝑟̅2 − 𝑟̅1 represents the directed line segment from 𝑄1 to 𝑄2 , as shown in the
Figure 14.

Figure 14: Coulomb vector force on point charges 𝑄1 and 𝑄2 ,

The vector 𝐹̅12 is the force in 𝑄2 and is shown for the case where 𝑄1 and 𝑄2 have the same sign. The
vector form of Coulomb’s law is
𝑄 𝑄
𝐹̅12 = 1 22 𝑎̅12
4𝜋𝜀𝑜 𝑅12
4

Where 𝑎̅12 is a unit vector in the direction of 𝑅12 , or


𝑅̅ 𝑟̅ −𝑟̅
𝑎̅12 = |𝑅̅12 | = |𝑟̅2 −𝑟̅1 | 5
12 2 1

The vector force in equation (4) can also be written as


𝑄1 𝑄2 (𝑟̅2 −𝑟̅1 )
𝐹̅12 = 6
4𝜋𝜀𝑜 |𝑟̅2 −𝑟̅1 |3

Example: A charge of 𝑄1 = 3 × 10−4 𝐶 is located at 𝑀(1,2,3) and a charge 𝑄2 = −10−4 𝐶 at 𝑀(2,0,5)


in a vacuum. Find the force exerted in 𝑄2 by 𝑄1 .

Solution
We shall make use of equations (4) and (5) to obtain the vector force.
𝑅̅12 = 𝑟̅2 − 𝑟̅1 = (2 − 1)𝑎̅𝑥 + (0 − 2)𝑎̅𝑦 + (5 − 3)𝑎̅𝑧

𝑅̅12 = 𝑎̅𝑥 − 2𝑎̅𝑦 + 2𝑎̅𝑧

|𝑅̅12 | = √12 + (−2)2 + 22 = √9 = 3

𝑅̅12 𝑎̅𝑥 − 2𝑎̅𝑦 + 2𝑎̅𝑧


𝑎̅12 = =
̅
|𝑅12 | 3

3 × 10−4 × (−10−4 ) 𝑎̅𝑥 − 2𝑎̅𝑦 + 2𝑎̅𝑧


𝐹̅12 = ×( )
1 3
4𝜋 × (36𝜋 × 10−9 ) × 32

𝑎̅𝑥 − 2𝑎̅𝑦 + 2𝑎̅𝑧


𝐹̅12 = −30 ( )𝑁
3
The magnitude of the force is 30𝑁 and the direction is specified by the unit vector which has been left
in parenthesis to display the magnitude of the force. The force in 𝑄2 may also be considered as those
components forces i.e.,
𝐹̅12 = −10𝑎̅𝑥 + 20𝑎̅𝑦 − 20𝑎̅𝑧 𝑁

The force expressed by Coulomb’s law is a mutual force for each of the two charges experiences a force
of the same magnitude, although of opposite direction. We might equally well have written
𝑄1 𝑄2 𝑄1 𝑄2
𝐹̅12 = −𝐹̅21 = 2 𝑎̅12 =− 2 𝑎̅21 6
4𝜋𝜀𝑜 𝑅12 4𝜋𝜀𝑜 𝑅21

Coulomb’s law is linear, for if we multiply 𝑄1 by a factor 𝑛, the force in 𝑄2 is also multiplied by the same
factor 𝑛.

Superposition Principle
It is also true that the force on a charge in the presence of several other charges is the sum of the forces
on that charge due to each of the other charges acting alone i.e., force on a charge 𝑄1 due to 𝑛 − 1
other charges 𝑄1 , 𝑄2 , … . , 𝑄𝑁 is the vector sum of the individual forces.

The principle states that if there are 𝑛 charges 𝑄1 , 𝑄2 , … . , 𝑄𝑁 located, respectively at 𝑟̅1 , 𝑟̅2 , … . , 𝑟̅𝑁 , the
resultant force 𝐹̅1𝑁 on a charge located at point 𝑟 is the vector sum of the forces exerted on charge 𝑄
by each of the charges 𝑄1 , 𝑄2 , … . , 𝑄𝑁 .
𝑄 𝑄 𝑄 𝑄 𝑄 𝑄
𝐹̅1𝑁 = 4𝜋𝜀1 𝑅22 𝑎̅12 + 4𝜋𝜀1 𝑅32 𝑎̅13 + ⋯ + 4𝜋𝜀1 𝑅𝑛2 𝑎̅1𝑁 7
𝑜 12 𝑜 13 𝑜 1𝑁

𝑄 𝑄𝑘 𝑄 𝑄𝑘 (𝑟̅1 −𝑟̅𝑁 )
𝐹̅1𝑁 = 4𝜋𝜀1 ∑𝑁 ̅1𝑁 = 4𝜋𝜀1 ∑𝑁
𝑛=1 𝑅2 𝑎 𝑛=1 |𝑟̅ −𝑟̅ |3 8
𝑜 1𝑁 𝑜 1 𝑁

Example: Point charges 1𝑚𝐶 and −2𝑚𝐶 are located at (3,2, −1) and (−1, −1,4), respectively.
Calculate the electric force on a 10𝑛𝐶 charge located at (0,3,1).

Solution
From Coulomb’s law,
𝑛
𝑄1 𝑄𝑘 (𝑟̅ − 𝑟̅𝑘 )
𝐹̅1𝑘 = ∑
4𝜋𝜀𝑜 |𝑟̅ − 𝑟̅𝑘 |3
𝑘=2
𝑟̅ − 𝑟̅1 = (0 − 3)𝑎̅𝑥 + (3 − 2)𝑎̅𝑦 + (1 − (−1))𝑎̅𝑧 = −3𝑎̅𝑥 + 𝑎̅𝑦 + 2𝑎̅𝑧 for 𝑄1 = 1𝑚𝐶

|𝑟̅ − 𝑟̅𝑘 | = √(−3)2 + 12 + 22 = √14

𝑟̅ − 𝑟̅1 = (0 − (−1))𝑎̅𝑥 + (3 − (−1))𝑎̅𝑦 + (1 − 4)𝑎̅𝑧 = 𝑎̅𝑥 + 4𝑎̅𝑦 − 3𝑎̅𝑧 for 𝑄2 = −2𝑚𝐶

|𝑟̅ − 𝑟̅𝑘 | = √12 + 42 + (−3)2 = √26

𝑄 𝑄1 (𝑟̅ − 𝑟̅1 ) 𝑄2 (𝑟̅ − 𝑟̅2 )


𝐹̅ = [ + ]
4𝜋𝜀𝑜 |𝑟̅ − 𝑟̅1 |3 |𝑟̅ − 𝑟̅2 |3

10 × 10−9 1 × 10−3 (−3𝑎̅𝑥 + 𝑎̅𝑦 + 2𝑎̅𝑧 ) (−2 × 10−3 )(𝑎̅𝑥 + 4𝑎̅𝑦 − 3𝑎̅𝑧 )
𝐹̅ = [ 3 + 3 ]
1
4𝜋 × ( × 10−9 ) (√14) (√26)
36𝜋
𝐹̅ = −6.507𝑎̅𝑥 − 3.817𝑎̅𝑦 + 7.506𝑎̅𝑧 𝑚𝑁

Exercise
1. A charge 𝑄𝐴 = −20𝜇𝐶 is located at 𝐴(−6,4,7) and a charge 𝑄𝐵 = 50𝜇𝐶 is located at 𝐵(5,8, −2)
in free space. If distances are given in meters, find:
a. 𝑅̅𝐴𝐵
b. 𝑅𝐴𝐵
c. The vector force exerted in 𝑄𝐴 by 𝑄𝐵 .

Electric Field Intensity


If we now consider one charge fixed in position say 𝑄1 and move a second charge slowly around, we
note that there exists everywhere a force on the second charge. In other words, the second charge is
displaying the existence of a force field. Let the second charge be a test charge 𝑄𝑡 .
The force on it is given by Coulomb’s law,
𝑄 𝑄
𝐹̅ = 4𝜋𝜀1 𝑅𝑡2 𝑎̅1𝑡 9
𝑜 1𝑡

Rating this force in a force per unit charge gives


𝐹̅ 𝑄1
= 2 𝑎̅1𝑡 10
𝑄𝑡 4𝜋𝜀𝑜 𝑅1𝑡

The quantity on the right side of equation (10) is a function of 𝑄1 and the direct line segment from 𝑄1
to the position of the test charge. This describes a vector field and is called the electric field intensity.
Electric field intensity is defined as the vector force on a unit positive test charge. It is measured in
Newtons per Coulomb (𝑁/𝐶) or Volts per meter (𝑉/𝑚).
Therefore, using a capital letter 𝐸̅ for electric field intensity, we have;
𝐹 ̅
𝐸̅ = 𝑄 11
𝑡

𝑄
𝐸̅ = 4𝜋𝜀 1𝑅2 𝑎̅1𝑡 12
𝑜 1𝑡

Equation (11) is the defining expression for electric field intensity, and equation (12) is the expression
for the electric field intensity due to a single point charge 𝑄1 in a vacuum. Let us disperse/neglect with
most of the subscripts in equation (12), reserving the right to use them again anytime, there is a
possibility of misunderstanding
𝑄
𝐸̅ = 4𝜋𝜀 1𝑅2 𝑎̅𝑅 13
𝑜

We should remember that 𝑅 is the magnitude of the vector 𝑅, the directed line segment from the point
at which the point charge 𝑄 is located to the point at which 𝐸̅ is desired, and 𝑎̅𝑅 is a unit vector in the
𝑅̅ direction.
Let us arbitrarily locate 𝑄1 at the centre of the spherical coordinate system. The unit vector 𝑎̅𝑅 then
becomes the radial unit vector 𝑎̅𝑟 , and 𝑅 is 𝑟. Hence,
𝑄
𝐸̅ = 4𝜋𝜀1 𝑟2 𝑎̅𝑟 14
𝑜

𝑄1
𝐸𝑟 = 15
4𝜋𝜀𝑜 𝑟 2

The field has a single radial component, and its inverse square law relationship is quite obvious.

Electric Field Intensity in Rectangular Coordinate


Writing these expressions in rectangular coordinate for a charge 𝑄 at the origin, we have
𝑅̅ = 𝑟̅ = 𝑥𝑎̅𝑥 + 𝑦𝑎̅𝑦 + 𝑧𝑎̅𝑧 16
and
(𝑥𝑎̅𝑥 +𝑦𝑎̅𝑦 +𝑧𝑎̅𝑧 )
𝑎̅𝑅 = 𝑎̅𝑟 = 17
√𝑥 2 +𝑦 2 +𝑧 2
Therefore,
𝑄 (𝑥𝑎̅ +𝑦𝑎̅ +𝑧𝑎̅ )
𝐸̅ = (𝑥 2 2 2 )
( 𝑥 2 𝑦2 2 𝑧 ) 18
4𝜋𝜀𝑜 +𝑦 +𝑧 √𝑥 +𝑦 +𝑧

𝑄 𝑥 𝑦 𝑧
𝐸̅ = 4𝜋𝜀 2 +𝑦 2 +𝑧 2 ) ( 𝑎̅𝑥 + 𝑎̅𝑦 + 𝑎̅𝑧 ) 19
𝑜 (𝑥 √𝑥 2 +𝑦 2 +𝑧 2 √𝑥 2 +𝑦 2 +𝑧 2 √𝑥 2 +𝑦 2 +𝑧 2

If a charge 𝑄 at the origin, in this case, the field no longer possess spherical symmetry (nor cylindrical
symmetry). Rectangular coordinate may also be used.
For a charge 𝑄 located at the same point 𝑟̅ ′ = 𝑥 ′ 𝑎̅𝑥 + 𝑦 ′ 𝑎̅𝑦 + 𝑧 ′ 𝑎̅𝑧 as shown in Figure 15.

Figure 15:
We find the field at a general field point 𝑟̅ ′ = 𝑥 ′ 𝑎̅𝑥 + 𝑦 ′ 𝑎̅𝑦 + 𝑧 ′ 𝑎̅𝑧 by expressing 𝑅̅ as 𝑟̅ − 𝑟̅ ′ and then
𝑄 𝑟̅−𝑟̅ ′
𝐸̅ (𝑟) = 4𝜋𝜀 ′ 2 × |𝑟̅−𝑟̅ ′ | 20
𝑜 |𝑟̅ −𝑟̅ |

𝑄(𝑟̅−𝑟̅ ) ′
𝐸̅ (𝑟) = 4𝜋𝜀 |𝑟̅−𝑟̅ ′ |3 21
𝑜

𝑄((𝑥−𝑥 ′ )𝑎̅𝑥 +(𝑦−𝑦 ′ )𝑎̅𝑦 +(𝑧−𝑧 ′ )𝑎̅𝑧 )


𝐸̅ (𝑟) = 3 22
4𝜋𝜀𝑜 ((𝑥−𝑥 ′ )2 +(𝑦−𝑦 ′ )2 +(𝑧−𝑧 ′ )2 ) ⁄2

Earlier, we defined a vector field as a vector function of a position vector, and this is emphasized by
letting 𝐸̅ be symbolized in functional notation by 𝐸̅ (𝑟).

Electric Field Intensity Due to Two Point Charge


Since the Coulomb force are linear, the electric field intensity due to two-point charge 𝑄1 at 𝑟̅1 and 𝑄2
at 𝑟̅2 is the sum of the forces on 𝑄𝑡 caused by 𝑄1 and 𝑄2 acting alone i.e.,
𝑄1 𝑄2
𝐸̅ (𝑟) = 4𝜋𝜀 2 𝑎̅1 + 4𝜋𝜀 2 𝑎̅2 23
𝑜 |𝑟̅ −𝑟̅1 | 𝑜 |𝑟̅ −𝑟̅2 |

Where 𝑎̅1 and 𝑎̅2 are unit vectors in the direction of 𝑟̅ − 𝑟̅1 and 𝑟̅ − 𝑟̅2 , respectively.
The vectors 𝑟̅ , 𝑟̅1 , 𝑟̅2 , 𝑟̅ − 𝑟̅1 , 𝑟̅ − 𝑟̅2 , 𝑎̅1 𝑎𝑛𝑑 𝑎̅2 are shown Figure 16.
Figure 16:

The vector addition of the total electric field intensity at 𝑃 due to 𝑄1 and 𝑄2 is made possible by the
linearity of Coulomb’s law.
If we add more charges at other positions, the field due to a point charge is
𝑄1 𝑄2 𝑄𝑛
𝐸̅𝑟 = 4𝜋𝜀 2 𝑎̅1 + 4𝜋𝜀 2 𝑎̅2 + ⋯ + 4𝜋𝜀 2 𝑎̅𝑛 24
𝑜 |𝑟̅ −𝑟̅1 | 𝑜 |𝑟̅ −𝑟̅2 | 𝑜 |𝑟̅ −𝑟̅𝑛 |

𝑄𝑚
𝐸̅𝑟 = ∑𝑛𝑚=0 4𝜋𝜀 2𝑎
̅𝑚 25
𝑜 |𝑟̅ −𝑟̅𝑚 |

Example: Given four identical 3 𝑛𝐶 charges located at 𝑃1 (1,1,0), 𝑃2 (−1,1,0), 𝑃3 (−1, −1,0) and
𝑃4 (1, −1,0) as shown. Find the electric field intensity 𝐸̅ at 𝑃(1,1,1).

Solution
𝑟̅ = 𝑎̅𝑥 + 𝑎̅𝑦 + 𝑎̅𝑧
𝑟̅1 = 𝑎̅𝑥 + 𝑎̅𝑦
𝑟̅2 = −𝑎̅𝑥 + 𝑎̅𝑦
𝑟̅3 = −𝑎̅𝑥 − 𝑎̅𝑦
𝑟̅4 = 𝑎̅𝑥 − 𝑎̅𝑦

|𝑟̅ − 𝑟̅1 | = √(1 − 1)2 + (1 − 1)2 + (1 − 0)2 = 1


|𝑟̅ − 𝑟̅2 | = √(1 − (−1))2 + (1 − 1)2 + (1 − 0)2 = √5
|𝑟̅ − 𝑟̅3 | = √(1 − (−1))2 + (1 − (−1))2 + (1 − 0)2 = √9 = 3
|𝑟̅ − 𝑟̅4 | = √(1 − 1)2 + (1 − (−1))2 + (1 − 0)2 = √5

(𝑎̅𝑥 −𝑎̅𝑥 )+(𝑎̅𝑦 −𝑎̅𝑦 )+(𝑎̅𝑧 −0)


𝑎̅1 = = 𝑎̅𝑧
1
(𝑎̅𝑥 −(−𝑎̅𝑥 ))+(𝑎̅𝑦 −𝑎̅𝑦 )+(𝑎̅𝑧 −0) 2𝑎̅ +𝑎̅
𝑎̅2 = 5
= 𝑥5 𝑧
√ √
(𝑎̅𝑥 −(−𝑎̅𝑥 ))+(𝑎̅𝑦 −(−𝑎̅𝑦 ))+(𝑎̅𝑧 −0) 2𝑎̅𝑥 +2𝑎̅𝑦 +𝑎̅𝑧
𝑎̅3 = 3
= 3
(𝑎̅𝑥 −𝑎̅𝑥 )+(𝑎̅𝑦 −(−𝑎̅𝑦 ))+(𝑎̅𝑧 −0) 2𝑎̅𝑦 +𝑎̅𝑧
𝑎̅4 = =
√5 √5

𝑄 3 × 10−9
= = 26.96𝑉
4𝜋𝜀𝑜 4𝜋 × 8.854 × 10−12
𝑄𝑚
Using 𝐸̅𝑟 = ∑𝑛𝑚=0 4𝜋𝜀 2 𝑎̅𝑚 ,
𝑜 |𝑟̅ −𝑟̅𝑚 |

1 1 1 1
𝐸̅𝑟 = 26.96 ( 𝑎
̅1 + 𝑎
̅ 2 + 𝑎
̅ 3 + 𝑎̅ )
|𝑟̅ − 𝑟̅1 |2 |𝑟̅ − 𝑟̅2 |2 |𝑟̅ − 𝑟̅3 |2 |𝑟̅ − 𝑟̅4 |2 4

1 1 2𝑎̅𝑥 + 𝑎̅𝑧 1 2𝑎̅𝑥 + 2𝑎̅𝑦 + 𝑎̅𝑧 1 2𝑎̅𝑦 + 𝑎̅𝑧


𝐸̅𝑟 = 26.96 ( 2
(𝑎̅𝑧 ) + 2 ( )+ 2( )+ 2( ))
1 √5 √5 3 3 √5 √5

2𝑎̅𝑥 + 𝑎̅𝑧 2𝑎̅𝑥 + 2𝑎̅𝑦 + 𝑎̅𝑧 2𝑎̅𝑦 + 𝑎̅𝑧


𝐸̅𝑟 = 26.96 (𝑎̅𝑧 + + + )
5√5 27 5√5
2 2 2 2 1 1 1
𝐸̅𝑟 = 26.96 (( + ) 𝑎̅𝑥 + ( + ) 𝑎̅𝑦 + (1 + + + ) 𝑎̅𝑧 )
5√5 27 27 5√5 5√5 27 5√5

𝐸̅𝑟 = 26.96 ((0.17 + 0.074)𝑎̅𝑥 + (0.074 + 0.17)𝑎̅𝑦 + (1 + 0.089 + 0.037 + 0.089)𝑎̅𝑧 )

𝐸̅𝑟 = 26.96(0.244𝑎̅𝑥 + 0.244𝑎̅𝑦 + 1.215𝑎̅𝑧 )

𝐸̅𝑟 = 6.58𝑎̅𝑥 + 6.58𝑎̅𝑦 + 32.75𝑎̅𝑧 𝑉/𝑚

Exercise
1. A charge of 0.3 𝜇𝐶 is located at 𝐴(25, −30,15) 𝑐𝑚 and a second charge of 0.5 𝜇𝐶 is at
𝐵(−10,8,12) 𝑐𝑚. Find the electric field intensity at:
a. the origin
b. 𝑃(15,20,50) 𝑐𝑚
2. Evaluate the sums
1+(−1)𝑚
a. ∑5𝑚=0 𝑚2 +1
4 (0.1)𝑚 +1
b. ∑𝑚=1
(4+𝑚2 )2

Field due to a Continuous Volume Charge Distribution


If we now visualize a region of space filled with a tremendous number of charges separated by minute
distance, the distribution of very small particles can be replaced with a smooth continuous distribution
described by a volume charge density.
We denote the volume charge density by 𝜌𝑣 having a unit of Coulomb per cubic meter (𝐶/𝑚3 ). The
small amount of charge ∆𝑄 in a small volume ∆𝑉 is
∆𝑄 = 𝜌𝑣 ∆𝑉
𝜌𝑣 is mathematically defined using a limiting process
∆𝑄
𝜌𝑣 = lim
∆𝑉→0 ∆𝑉

The total charge in the same finite volume is obtained by integrating throughout that volume

𝑄 = ∫ 𝜌𝑣 𝑑𝑣
𝑣𝑜𝑙

Field of a Line Charges


Up to this point, we have considered two types of charge distribution, their point charge and charge
distribution throughout a volume with a density 𝜌𝑣 𝐶/𝑚3. Let us assume a straight-line charge
extending along the z-axis in a cylindrical coordinate system from −∞ 𝑡𝑜 ∞ as shown Figure 17.
Figure 17:
We desire the electric field intensity 𝐸̅ at any and every point resulting from a uniform line charge 𝜌𝐿 .
A line charge of length 2𝑎 has a linear charge density 𝜌𝐿 (𝐶𝑚−1 ). The elemental length 𝑑𝑧 of the line
has a charge 𝜌𝐿 𝑑𝑧 producing an electric field 𝑑𝐸̅ at a point 𝑃 given by
𝜌𝐿 𝑑𝑧
𝑑𝐸̅ = 𝑎̅
4𝜋𝜀𝑜 (𝑟 + 𝑧 2 ) 𝑟
2

Where 𝜌𝐿 is the charge per unit length (𝐶𝑚−1 ). The field component in the direction 𝑟 (perpendicular
to the line) is
𝑟
𝑑𝐸̅𝑟 = 𝑑𝐸̅ cos 𝜃 = 𝑑𝐸̅
√(𝑟 2 + 𝑧 2 )
𝜌𝐿 𝑑𝑧 𝑟
𝑑𝐸̅𝑟 = 2 2
×
4𝜋𝜀𝑜 (𝑟 + 𝑧 ) √(𝑟 2 + 𝑧 2 )

Symmetry should always be considered first in order to determine two specific factors:
i. With which coordinates the field does not vary.
ii. Which components of the field are not present.
From symmetry, the positive and the negative 𝑧 components add to zero and the radial field at 𝑃 from
the line is then the integral of 𝑑𝐸̅𝑟 over the length 2𝑎 of the line.
𝜌𝐿 𝑟 +𝑎 𝑑𝑧
𝐸̅𝑟 = ∫
4𝜋𝜀𝑜 −𝑎 (𝑟 + 𝑧 2 )3⁄2
2

after the integration, using integration table or change of variable 𝑧 = 𝑟 cot 𝜃


𝑎
𝜌𝐿 𝑟 1 𝑧
𝐸̅𝑟 = ( )
4𝜋𝜀𝑜 𝑟 2 √𝑟 2 + 𝑧 2 −𝑎

Therefore, for a long line, (𝑎 >> 𝑟), the equation reduces to


𝜌𝐿
𝐸̅𝑟 = 𝑎̅
2𝜋𝜀𝑜 𝑟 𝑟

Field of a Sheet of Charge: Infinite Plane Charge


This is another basic charge configuration known as the infinite sheet of charge having a uniform density
of 𝜌𝑠 (𝐶/𝑚2 ). Such a charge distribution may often be used to approximate that found on the
conductors of a strip transmission line or a parallel plate capacitor. Static charge resides on conductor
surfaces and not in their interior. For this reason, 𝜌𝑠 is commonly known as surface charge density.
The charge distribution fairly now is complete - point, line, surface and volume or 𝑄, 𝜌𝐿 , 𝜌𝑠 and 𝜌𝑣 ,
respectively as shown in Figure 18.
Figure 18:
Now, let us develop an expression for 𝐸̅ due to charge uniformly distributed over an infinite plane with
density 𝜌𝑠 . The cylindrical coordinate system will be used with the charge in the 𝑧 = 0 plane as shown
in Figure 19.

Figure 19:

𝜌𝑠 𝑟𝑑𝑟𝑑𝜙 −𝑟𝑎̅𝑟 + 𝑧𝑎̅𝑧


𝑑𝐸̅ = 2 2 ( )
4𝜋𝜀𝑜 (𝑟 + 𝑧 ) √(𝑟 2 + 𝑧 2 )

Symmetry about the z-axis results in cancellation of the radial components.


−𝑟𝑎̅𝑟 + 𝑧𝑎̅𝑧
= 𝑎̅𝑧
√(𝑟 2 + 𝑧 2 )
Therefore,
𝜌𝑠 𝑟𝑑𝑟𝑑𝜙
𝑑𝐸̅ = 𝑎̅
4𝜋𝜀𝑜 (𝑟 2 + 𝑧 2 ) 𝑧
The field component in the direction z (perpendicular) to the surface is given as
𝑧
𝑑𝐸̅𝑟 = 𝑑𝐸̅ cos 𝜙 = 𝑑𝐸̅
√(𝑟 2 + 𝑧2)
2𝜋 ∞
𝜌𝑠 𝑟𝑑𝑟𝑑𝜙 𝑧
𝐸̅𝑟 = ∫ ∫ 2 2
0 0 4𝜋𝜀𝑜 (𝑟 )
+ 𝑧 √(𝑟 + 𝑧 2 )
2

2𝜋 ∞
𝜌𝑠 𝑟𝑧𝑑𝑟𝑑𝜙
𝐸̅𝑟 = ∫ ∫ 𝑎̅𝑧
0 0 4𝜋𝜀𝑜 (𝑟 2 + 𝑧 2 )3⁄2

𝜌𝑠 𝑧 −1
𝐸̅𝑟 = [ ] 𝑎̅𝑧
2𝜀𝑜 √(𝑟 2 + 𝑧 2 )
0
𝜌𝑠
𝐸̅𝑟 = 𝑎̅
2𝜀𝑜 𝑧
This result is for points above the 𝑥𝑦 plane. Below the 𝑥𝑦 plane the unit vector changes to −𝑎̅𝑧 . The
generalized form may be written using 𝑎̅𝑛 , the unit normal vector
𝜌𝑠
𝐸̅ = 𝑎̅
2𝜀𝑜 𝑛
The electric field is everywhere normal to the plane of the charge and its magnitude is independent of
the distance from the plane. For general volume charge distribution in free space is

𝜌𝑣 𝑑𝑣
𝐸̅ = ∫ 𝑎̅ 𝑓𝑟𝑒𝑒 𝑠𝑝𝑎𝑐𝑒 𝑜𝑛𝑙𝑦
2 𝑅
𝑣𝑜𝑙 4𝜋𝜀𝑜 𝑅

𝜌𝑣 𝑑𝑣
̅ = 𝜀𝑜 𝐸̅ = ∫
𝐷 𝑎̅
2 𝑅
𝑣𝑜𝑙 4𝜋𝑅

Gauss’s Law
̅𝑠 makes an angle 𝜃 with ∆𝑆
If we consider a closed surface, let ∆𝑆 be an incremental element 𝐷

̅𝑠 and ∆𝑆.
The flux crossing ∆𝑠 is the product of the normal component of 𝐷
∆𝑄 = 𝑓𝑙𝑢𝑥 𝑐𝑟𝑜𝑠𝑠𝑖𝑛𝑔 ∆𝑆 = 𝐷𝑠 𝑛𝑜𝑟𝑚𝑎𝑙 ∆𝑆
= 𝐷𝑠 cos 𝜃 ∆𝑆
̅𝑠 ∙ ∆𝑆
=𝐷
The total flux passing through the closed surface is obtained by adding the differential contributions
crossing each surface element ∆𝑆.

𝜓 = ∫ 𝑑𝜓 = ∮ ̅𝑠 ∙ 𝑑𝑆
𝐷
𝑐𝑙𝑜𝑠𝑒𝑑 𝑠𝑢𝑟𝑓.

The resultant integral is a closed surface integral and since the surface element 𝑑𝑆 always involves the
differential of two coordinate such as 𝑑𝑥𝑑𝑦, 𝜌𝑑𝜙𝑑𝜌 or 𝑟 2 sin 𝜃 𝑑𝜃𝑑𝜙. The integral is a double integral.
Any closed surface, real or imaginary may be called a Gaussian surface. Thus, the double or surface
integral with the circle (∮𝑠 ) implies a Gaussian or closed surface.

𝜓=∮ ̅𝑠 ∙ 𝑑𝑆 = 𝑐ℎ𝑎𝑟𝑔𝑒 𝑒𝑛𝑐𝑙𝑜𝑠𝑒𝑑 = 𝑄


𝐷
𝑐𝑙𝑜𝑠𝑒𝑑 𝑠𝑢𝑟𝑓.

When the charge enclosed several point charges in which case

𝑄 = ∑ 𝑄𝑛

Or a line charge 𝑄 = ∫ 𝜌𝑙 𝑑𝑙

Or a surface charge 𝑄 = ∫𝑆 𝜌𝑆 𝑑𝑆 (not necessary a closed surface)

Or a volume charge distribution 𝑄 = ∫𝑣𝑜𝑙 𝜌𝑣 𝑑𝑣

Gauss’s law may be written in terms of the charge distribution as


̅𝑠 ∙ 𝑑𝑆 = ∫ 𝜌𝑣 𝑑𝑣 = 𝜓
∮ 𝐷
𝑆
𝑣𝑜𝑙

this mathematical statement meaning simply that the total electric flux through any closed surface is equal to the
charge enclosed.

Example: using Gauss’s law proof that the radius 𝑟 = 𝑎 of a spherically enclosed surface has a point
charge 𝑄 at the origin.
̅𝑠 ∙ 𝑑𝑆 = 𝑄
∮ 𝐷
𝑆
Solution
𝑄 𝑄
̅𝑠 =
𝐷 𝑎̅𝑟 = 𝑎̅
4𝜋𝑟 2 4𝜋𝑎2 𝑟
𝑑𝑆 = 𝑟 2 sin 𝜃 𝑑𝜃 𝑑𝜙 𝑎̅𝑟 = 𝑎2 sin 𝜃 𝑑𝜃 𝑑𝜙 𝑎̅𝑟
𝑄
̅𝑠 ∙ 𝑑𝑆 =
𝐷 𝑎̅ × 𝑎2 sin 𝜃 𝑑𝜃 𝑑𝜙 𝑎̅𝑟
4𝜋𝑎2 𝑟
𝑄
̅𝑠 ∙ 𝑑𝑆 =
𝐷 sin 𝜃 𝑑𝜃 𝑑𝜙
4𝜋
𝜙=2𝜋 𝜃=𝜋 𝜙=2𝜋 𝜃=𝜋
𝑄
∫ ̅𝑠 ∙ 𝑑𝑆 = ∫
∫ 𝐷 ∫ sin 𝜃 𝑑𝜃 𝑑𝜙
4𝜋
𝜙=0 𝜃=0 𝜙=0 𝜃=0

𝜙=2𝜋 𝜃=𝜋
𝑄
= ∫ ∫ sin 𝜃 𝑑𝜃 𝑑𝜙
4𝜋
𝜙=0 𝜃=0

𝜙=2𝜋
𝑄
= ∫ (− cos 𝜃)𝜋0 𝑑𝜙
4𝜋
𝜙=0

𝜙=2𝜋
𝑄
= ∫ (− cos 𝜋 − (cos 0))𝑑𝜙
4𝜋
𝜙=0

𝜙=2𝜋 𝜙=2𝜋
𝑄 𝑄
= ∫ 2𝑑𝜙 = ∫ 𝑑𝜙
4𝜋 2𝜋
𝜙=0 𝜙=0

𝑄 𝑄
̅𝑠 ∙ 𝑑𝑆 =
∮ 𝐷 (𝜙)2𝜋
0 = (2𝜋 − 0) = 𝑄
𝑆 2𝜋 2𝜋

Example: A surface line charge infinite in extent with 𝜌𝐿 = 20 𝑛𝐶/𝑚 lies along the z-axis. Find electric
field 𝐸̅ at (6,8,3)𝑚.
Solution
In cylindrical coordinates,

𝑟 = √62 + 82 = 10 𝑚 the field is constant with z. Thus,


𝜌𝐿
𝐸̅ = 𝑎̅
2𝜋𝜀𝑜 𝑟 𝑟
20 × 10−9
𝐸̅ = −9 𝑎̅𝑟 = 36𝑎̅𝑟 𝑉/𝑚
2𝜋 × (10 ⁄36𝜋) × 10

Example: On the line described by 𝑥 = 2 𝑚, 𝑦 = −4 𝑚, there is a uniform charge distribution of density


𝜌𝐿 = 20𝑛𝐶/𝑚. Determine the electric field 𝐸̅ at (−2, −1,4)𝑚.
Solution
𝜌𝐿
𝐸̅ = 𝑎̅
2𝜋𝜀𝑜 𝑟 𝑟
The vector 𝑅̅ = (−2 − 2)𝑎̅𝑥 + (−1 − (−4))𝑎̅𝑦

𝑅̅ = −4𝑎̅𝑥 + 3𝑎̅𝑦

|𝑅̅ | = √(−4)2 + (3)2 = 5 𝑚


20 × 10−9 −4𝑎̅𝑥 + 3𝑎̅𝑦
𝐸̅ = −9 ×
2𝜋 × (10 ⁄36𝜋) × 5 5

𝐸̅ = −37.6𝑎̅𝑥 + 43.2𝑎̅𝑦 𝑉/𝑚

Example: Given the figure below, two uniform line charges of density 𝜌𝑙 = 4 𝑛𝐶/𝑚 lie in the 𝑥 = 0
plane at 𝑦 = ±4 𝑚. Find the electric field 𝐸̅ at (4,0,10)𝑚.

Solution
The magnitude of the field at 𝑃 would be
𝜌𝑙
𝐸=
2𝜋𝜀𝑜 𝑟
𝑟̅ = (4 − 0)𝑎̅𝑥 + (0 − (±4))𝑎̅𝑦 = 4𝑎̅𝑥 + 4𝑎̅𝑦

|𝑟̅ | = √(4)2 + (4)2 = 4√2 𝑚


4 × 10−9 18
𝐸= −9 = 𝑉/𝑚
2𝜋 × (10 ⁄36𝜋) × 4√2 √2

The field due to both line charges is superposition


18
𝐸̅𝑟 = 2 ( cos 45) 𝑎̅𝑥 = 18𝑎̅𝑥 𝑉/𝑚
√2

In an example of the evaluation of a volume integral, we shall find the total charge contained in a 2 𝑐𝑚
long of the electron beam shown in figure below.
Solution
From the illustration, we see that the charge density is
5
𝜌𝑣 = −5 × 10−6 𝑒 −10 𝜌𝑧 𝐶/𝑚3
The volume differential in cylindrical coordinates is given in the coordinate systems. Therefore,
0.04 2𝜋 0.01
5 𝜌𝑧
𝑄 = ∫ ∫ ∫ −5 × 10−6 𝑒 −10 𝜌𝑑𝜌𝑑𝜙𝑑𝑧
0.02 0 0

We integrate first with respect to 𝜙, since it is so easy


0.04 0.01
5 𝜌𝑧
𝑄 = ∫ ∫ −10−5 𝜋𝑒 −10 𝜌𝑑𝜌𝑑𝑧
0.02 0

And then with respect to 𝑧, because this will simplify the last integration with respect to 𝜌
0.01 0.04
−10−5 𝜋 −105 𝜌𝑧
𝑄=∫ [ 𝑒 𝜌𝑑𝜌]
−105 𝜌 0.02
0
0.01

𝑄 = ∫ −10−10 𝜋(𝑒 −2000𝜌 − 𝑒 −4000𝜌 ) 𝑑𝜌


0
Finally,
0.01
−10
𝑒 −2000𝜌 𝑒 −4000𝜌
𝑄 = −10 𝜋[ − ]
−2000 −4000 0

1 1 𝜋
𝑄 = −10−10 𝜋 ( − )=− = 0.0785 𝑝𝐶
2000 4000 40

ENERGY AND POTENTIAL


Energy Expended in Moving a Point Charge in an Electric Field
The electric field intensity was defined as the force on a unit test charge at that point we wish to find
the value of this vector field. If we attempt to move the test charge against the electric field, we have
to exert a force equal and opposite to that exerted by the field, and this requires us to expend energy
or do work. If we wish to move the charge in the direction of the field, our energy expenditure turns out
to be negative; we do not do the work.
Suppose we wish to move a charge 𝑄 a distance 𝑑𝐿̅ in an electric field 𝐸̅ , the force on 𝑄 arising from the
electric field is
𝐹̅𝐸 = 𝑄𝐸̅
where the subscript reminded us that this force arises from the field. The component of this force in the
direction 𝑑𝐿̅ which we must overcome is
𝐹𝐸𝐿 = 𝐹̅ ∙ 𝑎̅𝐿 = 𝑄𝐸̅ ∙ 𝑎̅𝐿
where 𝑎̅𝐿 is the unit vector in the direction of 𝑑𝐿̅.
The force which we must apply is equal and opposite to the force available with the field
𝐹𝑎𝑝𝑝𝑙𝑖𝑒𝑑 = −𝑄𝐸̅ ∙ 𝑎̅𝐿

and our expenditure of energy is the product of the force and distance, i.e., differential work done by
external source in moving 𝑄
𝑑𝑊 = −𝑄𝐸̅ ∙ 𝑎̅𝐿 𝑑𝐿 = −𝑄𝐸̅ ∙ 𝑑𝐿̅
where we have replaced 𝑎̅𝐿 𝑑𝐿 by a simpler expression 𝑑𝐿̅.
The work required to move the charge a finite distance must be determined by integrating work done
by electric field in moving a charge
𝑓𝑖𝑛𝑎𝑙

𝑊 = −𝑄 ∫ 𝐸̅ ∙ 𝑑𝐿̅
𝑖𝑛𝑖𝑡𝑖𝑎𝑙

where the path must be specified before the integral can be evaluated.

1
Example: Given the electric field 𝐸̅ = 2 (8𝑥𝑦𝑧𝑎̅𝑥 + 4𝑥 2 𝑧𝑎̅𝑦 − 4𝑥 2 𝑦𝑎̅𝑧 )𝑉/𝑚, find the differential
𝑧
amount of work done in moving a 6𝑛𝐶 charge a distance of 2𝜇𝑚, starting at 𝑃(2, −2,3) and proceeding
in the direction 𝑎̅𝐿 =
6 3 2
i. − 7 𝑎̅𝑥 + 7 𝑎̅𝑦 + 7 𝑎̅𝑧
6 3 2
ii. 𝑎̅ − 𝑎̅𝑦 − 𝑎̅𝑧
7 𝑥 7 7
3 6
iii. 𝑎̅
7 𝑥
+ 𝑎̅
7 𝑦

Solution
Using 𝑑𝑊 = −𝑄𝐸̅ ∙ 𝑎̅𝐿 𝑑𝐿
6 3 2
𝑑𝐿̅ = 𝑎̅𝐿 |𝑑𝐿| = (− 𝑎̅𝑥 + 𝑎̅𝑦 + 𝑎̅𝑧 ) × 2 × 10−6
7 7 7
12 6 4
𝑑𝐿̅ = (− 𝑎̅ + 𝑎̅ + 𝑎̅ ) × 10−6
7 𝑥 7 𝑦 7 𝑧
𝑑𝑊 = −𝑄𝐸̅ ∙ 𝑎̅𝐿 𝑑𝐿
1 12 6 4
𝑑𝑊 = −6 × 10−9 ( ̅
(8𝑥𝑦𝑧𝑎 𝑥 + 4𝑥 2
𝑧𝑎
̅ 𝑦 − 4𝑥 2
𝑦𝑎
̅ 𝑧 )) (− 𝑎
̅ 𝑥 + 𝑎
̅ 𝑦 + 𝑎̅ ) × 10−6
𝑧2 7 7 7 𝑧
96 24 16
𝑑𝑊 = −6 × 10−15 (− 𝑥𝑦𝑧 + 𝑥 2 𝑧 − 𝑥 2 𝑦)
7 7 7
96 24 16
𝑑𝑊𝑃(2,−2,3) = −6 × 10−15 (− (2 × −2 × 3) + (22 × 3) − (22 × −2))
7 7 7

𝑑𝑊𝑃(2,−2,3) = 1493 × 10−15 𝐽

The Line Integral


The integral expression for the work done in moving a point charge 𝑄 from one position to another. The
equation of work done is an example of a line integral, which in vector analysis notation always takes
the form of the integral along some presented path of the dot product of a vector field and a differential
vector path length 𝑑𝐿̅.
Without using vector analysis, we should have to write
𝑓𝑖𝑛𝑎𝑙

𝑊 = −𝑄 ∫ 𝐸̅𝐿 ∙ 𝑑𝐿̅
𝑖𝑛𝑖𝑡𝑖𝑎𝑙
where 𝐸̅𝐿 is te component of 𝐸̅ along 𝑑𝐿̅.
The procedure is indicated in the figure below, where a path has been chosen from an initial position 𝐵
to a final position 𝐴 and a uniform electric field selected for simplicity.
The path is divided into six segments ∆𝐿̅1 , ∆𝐿̅2 , … , ∆𝐿̅6 and the components of 𝐸̅ along each segment
are denoted by 𝐸𝐿1 , 𝐸𝐿2 , … , 𝐸𝐿6 .
The work involved in moving a charge 𝑄 from 𝐵 to 𝐴 is approximately
𝑊 = −𝑄(𝐸𝐿1 ∆𝐿1 + 𝐸𝐿2 ∆𝐿2 + ⋯ + 𝐸𝐿6 ∆𝐿6 )
or using vector notation,
𝑊 = −𝑄(𝐸̅1 ∙ ∆𝐿̅1 + 𝐸̅2 ∙ ∆𝐿̅1 + ⋯ + 𝐸̅6 ∙ ∆𝐿̅1 )

Graphical interpretation of a line integral in a uniform field.


and since we have assumed a uniform field.
𝐸̅1 = 𝐸̅2 = ⋯ = 𝐸̅6 = 𝐸̅
𝑊 = −𝑄𝐸̅ ∙ (∆𝐿̅1 + ∆𝐿̅1 + ⋯ + ∆𝐿̅1 )
The sum of vector segment is just the vector directed from the initial point 𝐵 to the point 𝐴, i.e., 𝐿̅𝐵𝐴 .
Therefore,
𝑊 = −𝑄𝐸̅ ∙ 𝐿̅𝐵𝐴 (𝑢𝑛𝑖𝑓𝑜𝑟𝑚 𝐸̅ )
This result for the uniform field can be obtained rapidly now from the integral expressions
𝐴

𝑊 = −𝑄 ∫ 𝐸̅ ∙ 𝑑𝐿̅
𝐵
as applied to a uniform field
𝐴

𝑊 = −𝑄𝐸̅ ∙ ∫ 𝑑𝐿̅
𝐵
where the last integral becomes 𝐿̅𝐵𝐴 and
𝑊 = −𝑄𝐸̅ ∙ 𝐿̅𝐵𝐴

Example: Given a non-uniform field 𝐸̅ = 𝑦𝑎̅𝑥 + 𝑥𝑎̅𝑦 + 2𝑎̅𝑧 , determine the work expended in carrying a
charge of 2𝐶 from 𝐵(1,0,1) to 𝐴(0.8,0.6,1) along the shorter arc of the circle 𝑥 2 + 𝑦 2 = 1, 𝑧 = 1.

Solution
We use
𝐴

𝑊 = −𝑄 ∫ 𝐸̅ ∙ 𝑑𝐿̅
𝐵
where 𝐸̅ is not necessarily constant. Working in rectangular coordinates, the differential path 𝑑𝐿̅ =
𝑑𝑥 ̅𝑎𝑥 + 𝑑𝑦 𝑎̅𝑦 + 𝑑𝑧 𝑎̅𝑧 and the integral becomes
𝐴

𝑊 = −2 ∫(𝑦𝑎̅𝑥 + 𝑥𝑎̅𝑦 + 2𝑎̅𝑧 ) ∙ (𝑑𝑥 ̅𝑎𝑥 + 𝑑𝑦 𝑎̅𝑦 + 𝑑𝑧 𝑎̅𝑧 )


𝐵

0.8 0.6 1

𝑊 = −2 ∫ 𝑦 𝑑𝑥 − 2 ∫ 𝑥 𝑑𝑦 + 4 ∫ 𝑑𝑧
1 0 1
The limits on the integrals have been chosen to agree with the initial and final values of the appropriate
variable of integration. Using the equation of the circular path
0.8 0.6

𝑊 = −2 ∫ √1 − 𝑥2 𝑑𝑥 − 2 ∫ √1 − 𝑦 2 𝑑𝑦 + 0
1 0

0.8 0.6
𝑊 = −2 [𝑥√1 − 𝑥 2 + sin−1 𝑥] − 2 [𝑦√1 − 𝑦 2 + sin−1 𝑦]
1 0

𝑊 = −2(0.48 + 0.927 − 0 − 1.571) − 2(0.48 + 0.644 − 0 − 0)


𝑊 = −0.96 𝐽
Example: Find the work expended in carrying 2𝐶 from 𝐵 to 𝐴 in the same field, but this time use the
straight-line path from 𝐵 to 𝐴.
Solution
We start by determining the equations of the straight line. Any two of the following three equations for
planes passing through the line are sufficient to define the line.
𝑦𝐴 − 𝑦𝐵
𝑦 − 𝑦𝐵 = (𝑥 − 𝑥𝐵 )
𝑥𝐴 − 𝑥𝐵
𝑧𝐴 − 𝑧𝐵
𝑧 − 𝑧𝐵 = (𝑦 − 𝑦𝐵 )
𝑦𝐴 − 𝑦𝐵
𝑥𝐴 − 𝑥𝐵
𝑥 − 𝑥𝐵 = (𝑧 − 𝑧𝐵 )
𝑧𝐴 − 𝑧𝐵
from the first equation, we have
𝑦 = −3(𝑥 − 1)
and from the second, we obtain
𝑧=1
thus,
0.8 0.6 1

𝑊 = −2 ∫ 𝑦 𝑑𝑥 − 2 ∫ 𝑥 𝑑𝑦 + 4 ∫ 𝑑𝑧
1 0 1
0.8 0.6
𝑦
𝑊 = −6 ∫ (𝑥 − 1) 𝑑𝑥 − 2 ∫ (1 − ⁄3) 𝑑𝑦
1 0
0.8 0.6
𝑥2 𝑦2
𝑊 = −6 [ + 𝑥] − 2 [𝑦 − ]
2 1
6 1
0.6
0.82 12 0.62 12
𝑊 = −6 [ + 0.8 − − 1] − 2 [0.6 − −1+ ]
2 2 6 6 1

𝑊 = −0.96 𝐽
Note that: the expressions for 𝑑𝐿̅ in the three coordinate systems use the differential lengths:
𝑑𝐿̅ = 𝑑𝑥 ̅𝑎𝑥 + 𝑑𝑦 𝑎̅𝑦 + 𝑑𝑧 𝑎̅𝑧

𝑑𝐿̅ = 𝑑𝜌 ̅𝑎𝜌 + 𝜌𝑑𝜙 𝑎̅𝜙 + 𝑑𝑧 𝑎̅𝑧

𝑑𝐿̅ = 𝑑𝑟 ̅𝑎𝑟 + 𝑟𝑑𝜃 𝑎̅𝜃 + 𝑟 sin 𝜃 𝑑𝜙 𝑎̅𝜙

The interrelationships among the several variables in each expression are determined from the specific
equations for the path. Thus, the potential difference between points at 𝜌 = 𝑎 and 𝜌 = 𝑏 is
𝑊 𝜌𝐿 𝑏
𝑉𝑎𝑏 = = ln
𝑄 2𝜋𝜀𝑜 𝑎
Potential difference between point 𝐴 and 𝐵 at radial distances 𝑟𝐴 and 𝑟𝐵 from a point charge 𝑄 is by
definition given from
𝑄
𝐸̅ = 𝐸𝑟 𝑎̅𝑟 = 𝑎̅
4𝜋𝜀𝑜 𝑟 2 𝑟
And 𝑑𝐿̅ = 𝑑𝑟 𝑎̅𝑟 , we have
𝐴

𝑉𝐴𝐵 = − ∫ 𝐸̅ ∙ 𝑑𝐿̅
𝐵
𝑟𝐴
𝑄
𝑉𝐴𝐵 = − ∫ 𝑎̅ ∙ 𝑑𝑟 𝑎̅𝑟
4𝜋𝜀𝑜 𝑟 2 𝑟
𝑟𝐵
𝑟𝐴
𝑄 𝑑𝑟 𝑄 1 𝑟𝐴
𝑉𝐴𝐵 =− ∫ 2 =− [ ]
4𝜋𝜀𝑜 𝑟 4𝜋𝜀𝑜 𝑟 𝑟𝐵
𝑟𝐵

𝑄 1 1
𝑉𝐴𝐵 = ( − )
4𝜋𝜀𝑜 𝑟𝐴 𝑟𝐵
If 𝑟𝐵 > 𝑟𝐴 , the potential difference 𝑉𝐴𝐵 is positive, indicating that energy is expended by the external
source in bringing the positive charge from 𝑟𝐵 to 𝑟𝐴 . Again, if the potential at point 𝐴 is 𝑉𝐴 and that at 𝐵
is 𝑉𝐵 , then
𝑉𝐴𝐵 = 𝑉𝐴 − 𝑉𝐵

The Potential Field of a Point Charge


We found an expression for the potential difference between two points located at 𝑟 = 𝑟𝐴 and 𝑟 = 𝑟𝐵
in the field of a point charge 𝑄 placed at the origin
𝑄 1 1
𝑉𝐴𝐵 = ( − ) = 𝑉𝐴 − 𝑉𝐵
4𝜋𝜀𝑜 𝑟𝐴 𝑟𝐵
How might we conveniently define a zero reference for potential? The simplest possibility is to let 𝑉 =
0 at infinity. If we let the point at 𝑟 = 𝑟𝐵 recede to infinity the potential at 𝑟𝐴 becomes
𝑄
𝑉𝐴 =
4𝜋𝜀𝑜 𝑟𝐴
or since there is no reason to identify this point with the 𝐴 subscript
𝑄
𝑉=
4𝜋𝜀𝑜 𝑟
The expression defines the potential at any point 𝑟 from a point charge 𝑄 at the origin, the potential at
infinite radius being taken as the zero reference. The potential is a scalar field and does not involve any
unit vector.

Equipotential Surface
Equipotential surface is a surface composed of all those points having the same values of potential. No
work is involved in moving a unit charge round on an equipotential surface for by definition, there is no
potential difference between any two points on this surface.

The Potential Field of a System of Charges


The potential field of a single point charge, which we shall identify as 𝑄1 and locate at 𝑟̅1 , involves only
the distance |𝑟̅ − 𝑟̅1 | from 𝑄1 to the point at 𝑟̅ where we are establishing the value of the potential.
For a zero reference at infinity, we have
𝑄
𝑉(𝑟̅ ) =
4𝜋𝜀𝑜 |𝑟̅ − 𝑟̅1 |
The potential arising from two charges, 𝑄1 at 𝑟̅1 and 𝑄2 at 𝑟̅2 , is a function only of |𝑟̅ − 𝑟̅1 | and |𝑟̅ − 𝑟̅2 |,
the distance from 𝑄1 and 𝑄2 to the field point, respectively,
𝑄 𝑄
𝑉(𝑟̅ ) = +
4𝜋𝜀𝑜 |𝑟̅ − 𝑟̅1 | 4𝜋𝜀𝑜 |𝑟̅ − 𝑟̅2 |
continuing to add charges, we find that the potential arising from 𝑛 point charges is
𝑄 𝑄 𝑄
𝑉(𝑟̅ ) = + +⋯+
|𝑟̅
4𝜋𝜀𝑜 − 𝑟̅1 | |𝑟̅
4𝜋𝜀𝑜 − 𝑟̅2 | 4𝜋𝜀𝑜 − 𝑟̅𝑛 |
|𝑟̅
or
𝑛
𝑄𝑚
𝑉(𝑟̅ ) = ∑
4𝜋𝜀𝑜 |𝑟̅ − 𝑟̅𝑚 |
𝑚=1

If each point charge is more represented as a small element of a continuous volume charge distribution
𝜌𝑣 ∆𝑣, then
𝜌𝑣 (𝑟̅ )∆𝑣1 𝜌𝑣 (𝑟̅ )∆𝑣2 𝜌𝑣 (𝑟̅ )∆𝑣𝑛
𝑉(𝑟̅ ) = + +⋯+
4𝜋𝜀𝑜 |𝑟̅ − 𝑟̅1 | 4𝜋𝜀𝑜 |𝑟̅ − 𝑟̅2 | 4𝜋𝜀𝑜 |𝑟̅ − 𝑟̅𝑛 |
when the number of elements becomes infinite, we obtain the integral expression

𝜌𝑣 (𝑟̅ ′ )𝑑𝑣 ′
𝑉(𝑟̅ ) = ∫
4𝜋𝜀𝑜 |𝑟̅ − 𝑟̅ ′ |
𝑣𝑜𝑙

where 𝑉(𝑟̅ ) is determined with respect to a zero-reference potential at infinity. 𝜌𝑣 (𝑟̅ ′ ) is the volume
charge density, 𝑑𝑣 ′ is the differential volume element combines to form a differential amount of charge
𝜌𝑣 (𝑟̅ ′ )𝑑𝑣 ′ located at 𝑟̅ ′ . The distance |𝑟̅ − 𝑟̅ ′ | is that distance from the source point to the field point.
The integral is a multiple (volume) integral. If the charge distribution takes the form of a line charge or
a surface charge, the integration is along the line or over the surface:
𝜌𝐿 (𝑟̅ ′ )𝑑𝐿′
𝑉(𝑟̅ ) = ∫
4𝜋𝜀𝑜 |𝑟̅ − 𝑟̅ ′ |

𝜌𝑠 (𝑟̅ ′ )𝑑𝑆 ′
𝑉(𝑟̅ ) = ∫
4𝜋𝜀𝑜 |𝑟̅ − 𝑟̅ ′ |
𝑆
Electric Flux Density
About 1837 the director of the Royal Society in London, Michael faraday became very interested in static
electric field and the effect of various insulating materials on these fields. He performed a lot of
experiments on the region between a pair of charged concentric spheres.

His experiments also showed that a large positive charge in the inner sphere induced a correspondingly
larger negative charge in the outer sphere, leading to a direct proportionality between the electric flux
and the charge on the inner sphere. If the electric flux is denoted by 𝜓 and the total charge in the inner
sphere by 𝑄, then from Faraday’s experiment
𝜓=𝑄
At the surface of the inner sphere, 𝜓 Coulomb’s of electric flux are produced by the charge 𝑄 (= 𝜓)
Coulombs distributed uniformly over a surface having an area of𝑚2 . The density of the flux at the
𝜓
surface ⁄ or 𝑄⁄ 𝐶/𝑚2 .
4𝜋𝑎2 4𝜋𝑎2
Therefore, electric flux density is the flux per area, it is measured in Coulombs per square meter.
Referring to the figure above, the electric flux density or displacement flux in the radial direction has a
value of
𝑄
̅𝑟=𝑎 =
𝐷 𝑎̅ (𝑖𝑛𝑛𝑒𝑟 𝑠𝑝ℎ𝑒𝑟𝑒)
4𝜋𝑎2 𝑟
𝑄
̅𝑟=𝑏 =
𝐷 𝑎̅ (𝑜𝑢𝑡𝑒𝑟 𝑠𝑝ℎ𝑒𝑟𝑒)
4𝜋𝑏 2 𝑟

And at a radial distance 𝑟, where 𝑎 ≤ 𝑟 ≤ 𝑏


𝑄
̅=
𝐷 𝑎̅
4𝜋𝑟 2 𝑟
i.e., inner sphere became smaller and smaller, while still retaining a charge of 𝑄, it becomes a point
charge in the limit. The result should be compared with the radial electric field intensity of a point charge
in free space
𝑄
𝐸̅ = 𝑎̅
4𝜋𝜀𝑜 𝑟 2 𝑟
In free space, therefore
̅ = 𝜀𝑜 𝐸̅
𝐷

Although the equation above is applicable only to a vacuum, it is not restricted solely to the field of a
point charge. For a general volume charge distribution in free space,

𝜌𝑣 𝑑𝑣
𝐸̅ = ∫ 𝑎̅ 𝑓𝑜𝑟 𝑓𝑟𝑒𝑒 𝑠𝑝𝑎𝑐𝑒
4𝜋𝜀𝑜 𝑅 2 𝑅
𝑣𝑜𝑙

where this relationship was developed from the field of a single point charge. In a similar manner, the
electric flux density leads to
𝜌𝑣 𝑑𝑣
̅= ∫
𝐷 𝑎̅ 𝑓𝑜𝑟 𝑓𝑟𝑒𝑒 𝑠𝑝𝑎𝑐𝑒
4𝜋𝑅 2 𝑅
𝑣𝑜𝑙

̅ in rectangular coordinates at 𝑃(2, −3,6) produced by:


Example: Calculate 𝐷
(a) A point charge 𝑄 = 55 𝑚𝐶 at 𝑄(−2,3, −6)
(b) A uniform line charge 𝜌𝐿 = 20 𝑚𝐶/𝑚 on the x-axis i.e., −∞ ≤ 𝑥 ≤ ∞
(c) A uniform surface charge density 𝜌𝑆 = 120 𝜇𝐶/𝑚2 on the plane 𝑧 = 5 𝑚.

Solution
(a)
𝑄 𝑅̅𝑃𝑄
̅ = 𝜀𝑜 𝐸̅ =
𝐷 2
4𝜋|𝑅̅𝑃𝑄 | |𝑅̅𝑃𝑄 |

𝑅̅𝑃𝑄 = (2 − (−2))𝑎̅𝑥 + (−3 − 3)𝑎̅𝑦 + (6 − (−6))𝑎̅𝑧 = 4𝑎̅𝑥 − 6𝑎̅𝑦 + 12𝑎̅𝑧

|𝑅̅𝑃𝑄 | = √42 + (−62 ) + 122 = 14

55 × 10−3 × (4𝑎̅𝑥 − 6𝑎̅𝑦 + 12𝑎̅𝑧 )


̅=
𝐷
4𝜋 × 143

̅ = 6.38𝑎̅𝑥 − 9.57𝑎̅𝑦 + 19.14𝑎̅𝑧 𝜇𝐶/𝑚2


𝐷

(b)
𝜌𝑙 𝜌𝑙 𝑅̅𝑃𝑋
𝐸̅ = 𝑎̅𝑟 =
2𝜋𝜀𝑜 𝑟 2𝜋𝜀𝑜 |𝑅̅𝑃𝑋 | |𝑅̅𝑃𝑋 |

𝑅̅𝑃𝑋 = (2 − 𝑥)𝑎̅𝑥 + (−3 − 0)𝑎̅𝑦 + (6 − 0)𝑎̅𝑧 = (2 − 𝑥)𝑎̅𝑥 − 3𝑎̅𝑦 + 6𝑎̅𝑧

since the infinite line charge is along X axis so the E field at point P is having only y and z components
present and the x component is cancelled due to symmetry so
𝑅̅𝑃𝑋 = −3𝑎̅𝑦 + 6𝑎̅𝑧

|𝑅̅𝑃𝑋 | = √(−32 ) + 62 = √45


𝜌𝑙 𝑅̅𝑃𝑋
̅ = 𝜀𝑜 𝐸̅ =
𝐷
2𝜋|𝑅̅𝑃𝑋 | |𝑅̅𝑃𝑋 |

20 × 10−3 × (−3𝑎̅𝑦 + 6𝑎̅𝑧 )


̅=
𝐷 2
2𝜋(√45)

̅ = −212 𝑎̅𝑦 + 424 𝑎̅𝑧 𝜇𝐶/𝑚2


𝐷

(c)
𝜌𝑆
𝐸̅ = 𝑎̅
2𝜀𝑜 𝑛
for infinite surface charge density also 𝑧 = 5 is an infinite x-y plane located at 𝑧 = 5 and
the charge is spread on this plane, so
𝜌 𝑅̅
̅ = 𝜀𝑜 𝐸̅ = 𝑆 𝑃𝑍
𝐷
2 |𝑅̅𝑃𝑍 |
120 × 10−6
̅=
𝐷 = 60 𝑎̅𝑧 𝜇𝐶/𝑚2
2

Divergence and Maxwell Equation


̅ over a
According to Gauss’s law, the integral of the normal component of the electric flux density 𝐷
closed surface yields the electric charge 𝑄 enclosed thus,

̅𝑆 ∙ 𝑑𝑆 = ∭ 𝜌𝑑𝑣 = 𝑄
∯𝐷

For incremental charges ∆𝑄 and volumes ∆𝑉,

∯ 𝐷𝑛 ∙ 𝑑𝑆 = 𝜌∆𝑉 = ∆𝑄

Where 𝐷𝑛 is the electric flux density normal to the surface.


If the charge density is not uniform throughout ∆𝑉, we have in the limit as the volume ∆𝑉 shrinks to
zero that

∬ 𝐷𝑛 ∙ 𝑑𝑆
lim ̅=𝜌
= 𝑑𝑖𝑣 𝐷
∆𝑉→0 ∆𝑉
Where 𝑑𝑖𝑣 𝐷 ̅ is the divergence of 𝐷
̅ = 𝜌. The divergence of 𝐷
̅ yields the electric charge density 𝜌 at a
̅
point. 𝑑𝑖𝑣 𝐷 has a value wherever charge is present.
̅ is the outflow of flux from a small
Physical interpretation: The divergence of the vector flux density 𝐷
closed surface per unit volume as the volume shrinks to zero.
̅ is expressed as the scalar or dot product of the 𝑑𝑒𝑙 operator
In vector notation, the divergence of 𝐷
̅
∇ 𝑎𝑛𝑑 𝐷. Thus,
𝛿 𝛿 𝛿
̅ =∇∙𝐷
𝑑𝑖𝑣 𝐷 ̅=( 𝑥̂ + 𝑦̂ + 𝑧̂ ) ∙ (𝑥̂ 𝐷𝑥 + 𝑦̂ 𝐷𝑦 + 𝑧̂ 𝐷𝑧 ) = 𝜌
𝛿𝑥 𝛿𝑦 𝛿𝑧
Taking the dot product yield
𝛿𝐷𝑥 𝛿𝐷𝑦 𝛿𝐷𝑧
̅=
𝑑𝑖𝑣 𝐷 + + = 𝜌 𝑟𝑒𝑐𝑡𝑎𝑛𝑔𝑢𝑙𝑎𝑟
𝛿𝑥 𝛿𝑦 𝛿𝑧
Where 𝑥̂ ∙ 𝑥̂ = 1, 𝑦̂ ∙ 𝑦̂ = 1 𝑎𝑛𝑑 𝑧̂ ∙ 𝑧̂ = 1
1 𝛿 1 𝛿𝐷ϕ 𝛿𝐷𝑧
̅=
𝑑𝑖𝑣 𝐷 (𝜌 𝐷𝜌 ) + + = 𝜌 𝑐𝑦𝑙𝑖𝑛𝑑𝑟𝑖𝑐𝑎𝑙
𝜌 𝛿𝜌 𝜌 𝛿𝜙 𝛿𝑧
1 𝛿 2 1 𝛿 1 𝛿𝐷ϕ
̅=
𝑑𝑖𝑣 𝐷 (𝑟 𝐷𝑟 ) + (sin 𝜃 𝐷θ ) + = 𝜌 𝑠𝑝ℎ𝑒𝑟𝑖𝑐𝑎𝑙
2
𝑟 𝛿𝜌 𝑟 sin 𝜃 𝛿𝜃 𝑟 sin 𝜃 𝛿𝜙

̅ field.
Exercise: Determine an expression for the volume charge density associated with each 𝐷
2 2
(a) ̅ = 4𝑥𝑦 𝑎𝑥 + 2𝑥 𝑎𝑦 − 2𝑥2 𝑦 𝑎𝑧
𝐷
𝑧 𝑧 𝑧
(b) ̅ = 𝑧 sin 𝜙 𝑎𝜌 + 𝑧 cos 𝜙 𝑎𝜙 + 𝜌 sin 𝜙 𝑎𝑧
𝐷
(c) ̅ = sin 𝜃 sin 𝜙 𝑎𝑟 + cos 𝜃 sin 𝜙 𝑎𝜃 + cos 𝜙 𝑎𝜙
𝐷
Maxwell’s First Equation
For any continuously differentiable vector 𝐴̅, we have

∮ 𝐴̅ ∙ 𝑑𝑆̅ = ∫ ∇ ∙ 𝐴̅ 𝑑𝑣
𝑣
This is known as divergence theorem. It has been stated earlier that

̅ ∙ 𝑑𝑆̅ = ∮ 𝜌𝑣 𝑑𝑣
𝜓=∮ 𝐷 𝐺𝑎𝑢𝑠𝑠 ′ 𝑠 𝑙𝑎𝑤
𝑠 𝑣
Applying Gauss’s law, we have

̅ ∙ 𝑑𝑆̅ = ∫ ∇ ∙ 𝐷
∮ 𝐷 ̅ 𝑑𝑣 = ∮ 𝜌𝑣 𝑑𝑣
𝑠 𝑣 𝑣

̅ = 𝜌𝑣 is called maxwell’s first equation in


This is maxwell’s first equation in integral form, but ∇ ∙ 𝐷
differential form.

The Equation of Poisson’s and Laplace


̅ = 𝜌𝑣 (Maxwell’s 1st equation in differential form)
We know that ∇ ∙ 𝐷
∇ ∙ 𝜀𝑜 𝐸̅ = 𝜌𝑣
𝜌𝑣
∇ ∙ 𝐸̅ =
𝜀𝑜
Replacing 𝐸̅ = −∇V we have
𝜌𝑣
∇ ∙ ∇V = −
𝜀𝑜
𝜌𝑣
∇2 V = −
𝜀𝑜

this is called Poisson’s equation which relate the volume charge density 𝜌𝑣 at a given point to the second
derivative of V in the region of that point.
In a region where the charge density 𝜌𝑣 is zero i.e., ∇2 V = 0, this is the Laplace equation.
The general problem of finding V in the field of aa given charge distribution amount to finding a solution
of either Laplace or Poisson equation that will satisfy the given boundary conditions.

Boundary Conditions for Perfect Dielectric Materials


When there are two different dielectrics, or a dielectric and a conductor, then boundary condition
occurs. This is the condition at the surface of a conductor whereby the tangential fields are zero and the
normal electric flux density is equal to the surface charge density on the conductor.
Let us first consider the interface between two dielectrics having permittivitives 𝜀1 and 𝜀2 and occupying
region 1 and 2 as shown below.

The boundary between perfect dielectric of permittivitives 𝜀1 and 𝜀2 is the tangential components is
obtained using
∮ 𝐸̅ ∙ 𝑑𝐿̅ = 0

Around the small closed path in the left, obtaining


𝐸tan 1 ∆𝑤 − 𝐸tan 2 ∆𝑤 = 0
Therefore, as ∆ℎ becomes negligible, the normal component of 𝐸̅ to becomes negligible and
𝐸tan 1 ∆𝑤 = 𝐸tan 2 ∆𝑤
̅ is
If the tangential electric field intensity is continuous across the boundary, then tangential 𝐷
discontinuous for
𝐷tan 1 𝐷tan 2
= 𝐸tan 1 = 𝐸tan 2 =
𝜀1 𝜀2
Or
𝐷tan 1 𝜀1
=
𝐷tan 2 𝜀2

The boundary conditions on the normal components are found by applying Gauss’s law, the surface
integral of 𝐷̅ over a closed surface equals the charge enclosed. The flux leaving the top and bottom
surface is the difference
𝐷𝑛1 ∆𝑆 − 𝐷𝑛2 ∆𝑆 = ∆𝑄 = 𝜌𝑆 ∆𝑆
From which
𝐷𝑛1 − 𝐷𝑛2 = 𝜌𝑆
When the charge density is zero in the interface
𝐷𝑛1 = 𝐷𝑛2
̅ is continuous. It follows that
Or the normal component of 𝐷
𝜀1 𝐸𝑛1 = 𝜀2 𝐸𝑛2

̅ are continuous
The refraction of 𝐷 at a dielectric interface since the normal components of 𝐷
𝐷𝑛1 = 𝐷1 cos 𝜃1 = 𝐷𝑛2 = 𝐷2 cos 𝜃2
The ratio of the tangential components is given by
𝐷tan 1 𝐷1 sin 𝜃1 𝜀1
= =
𝐷tan 2 𝐷2 sin 𝜃2 𝜀2
Or
𝜀2 𝐷1 sin 𝜃1 = 𝜀1 𝐷2 sin 𝜃2
And the division of the equation by
tann 𝜃1 𝜀1
=
tan 𝜃2 𝜀2
̅ in
In the above diagram, we have assumed that 𝜀1 > 𝜀2 and therefore, 𝜃1 > 𝜃2 . The magnitude of 𝐷
region 2 may be found as

𝜀2 2
𝐷2 = 𝐷1 √cos2 𝜃1 + ( ) sin2 𝜃2
𝜀1
And the magnitude of 𝐸2 is
𝜀1 2
𝐸2 = 𝐸1 √sin2 𝜃1 + ( ) cos2 𝜃2
𝜀2
An inspection of these equations shows that 𝐷 is larger in the region of larger permittivity (unless 𝜃1 = 𝜃2 = 0◦
where the magnitude is unchanged) and that 𝐸 is larger in the region of smaller permittivity (unless 𝜃1 = 𝜃2 =
90° , where its magnitude is unchanged).

Electric Potential and its Gradient


Since the work per charge is the electric potential difference ∆𝑉 between the point 𝑎 and 𝑏
∆𝑉 = 𝐸 ∆𝑥 (𝐽/𝐶 = 𝑉)

∆𝑉
Therefore, electric field 𝐸 = (𝑉/𝑚)
∆𝑥
In vector notation
∆𝑉
𝐸̅ = 𝑥𝐸 = −𝑥̅ (𝑉/𝑚)
∆𝑥
Negative sign reminds us that moving against the electric field results in positive work. When expanding
to the general case where 𝐸̅ also has 𝑦 and 𝑧 components give
𝛿 𝛿 𝛿
𝐸̅ = − (𝑥̅ + 𝑦̅ + 𝑧̅ ) 𝑉 (𝑉/𝑚)
𝛿𝑥 𝛿𝑦 𝛿𝑧
Therefore, 𝐸̅ is called the gradient of 𝑉. Abbreviating gradient to 𝑔𝑟𝑎𝑑 or the operator 𝑑𝑒𝑙 (∇) gives
𝐸̅ = −𝑔𝑟𝑎𝑑 𝑉 = −∇𝑉 (𝑉/𝑚)

The Electric Dipole


The electric dipole consists of the charges, a positive and a negative charge of the same magnitude,
separated by a distance 𝑑, which is small compared to the distance 𝑟 to the point 𝑃 at which we require
the electric potential 𝑉 and the electric field intensity 𝐸̅ . The dipole is shown in the figure below.
The distant point 𝑃 is described by the spherical coordinates 𝑟, 𝜃 𝑎𝑛𝑑 𝜙 = 90° , in view of azimuthal
symmetry. The positive and negative point charges have separation 𝑑 and rectangular coordinates
(0. 0, 𝑑⁄2) and (0. 0, −𝑑⁄2), respectively.
The above arrangement of two charges +𝑄 and −𝑄 forms a dipole. The electric potential at 𝑃 is the
sum of the potentials due to the individual charges. Let the distances from a dipole. The electric
potential at 𝑃 is the sum of the potentials due to the individual charges. Let the distance from +𝑄 and
−𝑄 to 𝑃 be 𝑅1 and 𝑅2 , respectively, and write the total potential as
𝑄 1 1 𝑄 𝑅2 − 𝑅1
𝑉= ( − )=
4𝜋𝜀𝑜 𝑅1 𝑅2 4𝜋𝜀𝑜 𝑅1 𝑅2
Note that: the plane 𝑧 = 0, midway between the two-point charges is the locus of points for which 𝑅1 =
𝑅2 and is therefore at zero potential, as are all points infinity

For a distant point 𝑃, 𝑅1 is essentially parallel to 𝑅2 and we find that 𝑅2 − 𝑅1 = 𝑑 cos 𝜃, therefore,
𝑅2 − 𝑅1 = 𝑑 cos 𝜃
The final result is then
𝑄 𝑑 cos 𝜃
𝑉=
4𝜋𝜀𝑜 𝑟 2
Again, we note that the plane 𝑧 = 0 (𝜃 = 90° ) is at zero potential. Using the gradient relationship in
spherical coordinates
𝛿𝑉 1 𝛿𝑉 1 𝛿𝑉
𝐸̅ = −∇𝑉 = − ( 𝑎̅𝑟 + 𝑎̅𝜃 + 𝑎̅ )
𝛿𝑟 𝑟 𝛿𝜃 𝑟 sin 𝜃 𝛿𝜙 𝜙
We obtain
𝑄𝑑 cos 𝜃 𝑄𝑑 sin 𝜃
𝐸̅ = − (− 𝑎̅𝑟 − 𝑎̅ )
2𝜋𝜀𝑜 𝑟 3 4𝜋𝜀𝑜 𝑟 3 𝜃
Or
𝑄𝑑
𝐸̅ = (2 cos 𝜃 𝑎̅𝑟 + sin 𝜃 𝑎̅𝜃 )
2𝜋𝜀𝑜 𝑟 3
The potential field of the dipole 𝑉 may be simplified by making use of the dipole moment. If the vector
length from +𝑄 and −𝑄 is 𝑑̅, then the dipole moment is defined as 𝑄𝑑̅, it is denoted by symbol 𝑃̅, Thus,
𝑃̅ = 𝑄𝑑̅
The unit of 𝑃̅ are 𝐶. 𝑚. since 𝑑̅ ∙ 𝑎̅𝑟 = 𝑑 cos 𝜃, then, we have
𝑃̅ ∙ 𝑎̅𝑟
𝑉=
4𝜋𝜀𝑜 𝑟 2
This result may be generated as
1 𝑟̅ − 𝑟̅ ′
𝑉= ̅∙
𝑃
4𝜋𝜀𝑜 |𝑟̅ − 𝑟̅ ′ |2 |𝑟̅ − 𝑟̅ ′ |
Where 𝑟̅ locates the field point 𝑃 and 𝑟̅ ′ determines the dipole centre.
Example: An electric dipole located at the origin in free space has a moment 𝑃̅ = 3𝑎̅𝑥 − 2𝑎̅𝑦 + 𝑎̅𝑧 𝑛𝐶 ∙
𝑚. Find the potential 𝑉 at 𝑃𝐴 (2, 3, 4).
Solution
𝑃̅ ∙ 𝑎̅𝑟
𝑉=
4𝜋𝜀𝑜 𝑟 2

𝑟̅ = 2𝑎̅𝑥 + 3𝑎̅𝑦 + 4𝑎̅𝑧

2𝑎̅𝑥 + 3𝑎̅𝑦 + 4𝑎̅𝑧 2𝑎̅𝑥 + 3𝑎̅𝑦 + 4𝑎̅𝑧


𝑎̅𝑟 = =
√22 + 32 + 42 √29

3𝑎̅𝑥 − 2𝑎̅𝑦 + 𝑎̅𝑧 2𝑎̅𝑥 + 3𝑎̅𝑦 + 4𝑎̅𝑧


𝑉=( ) × 10−9 × ( )
4𝜋𝜀𝑜 × 29 √29

(6 + 6 + 4) × 10−9
𝑉= = 0.23 𝑉
10−9
4𝜋 × ( 36𝜋 ) × 29 × 5.385

The Method of Images


One important characteristic of the dipole field is the infinite plane at zero potential that exists midway
between the two charges. Such a plane may be represented by a vanishingly thin conducting plane that
is infinite in extent. The conductor is an equipotential surface at a potential 𝑉 = 0, and the electric field
intensity is therefore normal to the surface.
Thus, if we replace the dipole configuration shown in Figure (a) below with the single charge and
conducting plane shown in Figure (b), the fields in the upper half of each figure are the same. Below the
conducting plane, all fields are zero, as we have not provided any charges in that region. We might also
substitute a single negative charge below the conducting plane for the dipole arrangement and obtain
equivalence for the fields in the lower half of each the region. If we approach this equivalence from the
opposite point of view, we begin with a single charge above a perfectly conducting plane and then see
that we may maintain the same fields above the plane by removing the plane and locating a negative
charge at a symmetrical location below the plane. This charge is called the image of the original charge,
and it is the negative of that value.

If we can do this once, linearity allows us to do it again and again, and thus any charge configuration
above an infinite ground plane may be replaced by an arrangement composed of the given charge
configuration, its image, and no conducting plane. This is suggested by the two illustrations of figure
below. In many cases, the potential field of the new system is much easier to find since it does not
contain the conducting plane with its unknown surface charge distribution.

Example: Find the surface charge density at 𝑃(2, 5, 0) on the conducting plane 𝑧 = 0 if there is a line
charge of 30 𝑛𝐶/𝑚 located at 𝑥 = 0, 𝑧 = 3, as shown in Figure (a) below. We remove the plane and
install an image line charge of −30 𝑛𝐶/𝑚 at 𝑥 = 0, 𝑧 = −3, as illustrated in Figure (b). The field at 𝑃
may now be obtained by superposition of the known fields of the line charges.

Solution
The radial vector from the positive line charge to 𝑃 is 𝑅+ = 2𝑎̅𝑥 − 3𝑎̅𝑧 , while 𝑅− = 2𝑎̅𝑥 + 3𝑎̅𝑧 ,. Thus,
the individual fields are
𝜌𝐿
𝐸+ = 𝑎̅
2𝜋𝜀𝑜 𝑅+ 𝑅+

2𝑎̅𝑥 − 3𝑎̅𝑧 2𝑎̅𝑥 − 3𝑎̅𝑧


𝑎̅𝑅+ = =
√22 + (−3)2 √13
30 × 10−9 2𝑎̅𝑥 − 3𝑎̅𝑧
𝐸+ = −9 ×
10 √13
2𝜋 × ( 36𝜋 ) × √13
and
−30 × 10−9 2𝑎̅𝑥 + 3𝑎̅𝑧
𝐸− = −9 ×
10 √13
2𝜋 × ( 36𝜋 ) × √13

Adding these results, we have


−180 × 10−9 𝑎̅𝑧
𝐸 = 𝐸+ + 𝐸− = = −249 𝑎̅𝑧 𝑉/𝑚
10−9
2𝜋 × ( 36𝜋 ) × 13

This then is the field at (or just above) 𝑃 in both the configurations of figure above, and it is certainly
satisfying to note that the field is normal to the conducting plane, as it must be. Thus, 𝐷 = 𝜀𝑜 𝐸 =
−2.20 𝑎̅𝑧 𝑛𝐶/𝑚2 , and because this is directed toward the conducting plane, 𝜌𝑆 is negative and has a
value of −2.20 𝑎̅𝑧 𝑛𝐶/𝑚2 at 𝑃.

Exercise:
A perfectly conducting plane is located in free space at 𝑥 = 4, and a uniform infinite line charge of
40 𝑛𝐶/𝑚 lies along the line 𝑥 = 6, 𝑦 = 3. Let 𝑉 = 0 at the conducting plane. At 𝑃(7, −1, 5).
Find: (a) 𝑉 ;
(b) 𝐸.

Capacitors and Capacitance


Capacitor is a device for storing electrical charges and hence, electrical energy. Imagine a finite
conductor situated a long distance
Example: Find the capacitance of a parallel plate capacitor containing dielectrics 𝜀𝑟1 = 1.5 and 𝜀𝑟2 =
3.5, each comprising one-half the volume as shown in the figure below. Here 𝐴 = 2𝑚2 and 𝑑 = 10−3 𝑚.

Solution
𝜀𝑜 𝜀𝑟1 𝐴1 (8.854 × 10−12 ) × 1.5 × 1
𝐶1 = = = 13.3 𝑛𝐹
𝑑 10−3
Similarly,
𝜀𝑜 𝜀𝑟2 𝐴2 (8.854 × 10−12 ) × 3.5 × 1
𝐶2 = = = 31.0 𝑛𝐹
𝑑 10−3
Then, 𝐶 = 𝐶1 + 𝐶2 = 13.3 𝑛𝐹 + 31.0 𝑛𝐹 = 44.3 𝑛𝐹

Example: Repeat the example above if the two dielectrics each occupy one-half of the space between
the plates but the interface is parallel to the plates.
Solution
𝜀𝑜 𝜀𝑟1 𝐴 (8.854 × 10−12 ) × 1.5 × 2
𝐶1 = = = 53.1 𝑛𝐹
𝑑1 10−3⁄
2
Since 𝑑1 = 𝑑2 = 𝑑⁄2

Similarly,
𝜀𝑜 𝜀𝑟1 𝐴 (8.854 × 10−12 ) × 3.5 × 2
𝐶2 = = = 124 𝑛𝐹
𝑑2 10−3⁄
2
𝐶1 𝐶2
Then, 𝐶 = 𝐶 = 37.2 𝑛𝐹
1 + 𝐶2

Example: Find the voltage across each dielectric in the capacitors shown in the figure below, when the
applied voltage is 200 𝑉, 𝜀𝑟1 = 5, 𝜀𝑟2 = 1, 𝑑1 = 1 𝑚𝑚, 𝑑2 = 3 𝑚𝑚 and 𝐴 = 1 𝑚2.

Solution
𝜀𝑜 × 5 × 1
𝐶1 = = 5000 𝜀𝑜
10−3
𝜀𝑜 × 1 × 1 1000 𝜀𝑜
𝐶2 = =
3 × 10−3 3
𝐶1 𝐶2
And 𝐶 = = 312.5 𝜀𝑜 𝑛𝐹 = 2.77 𝑛𝐹
𝐶1 + 𝐶2

̅ field within the capacitor is now found from


The 𝐷
𝑄 𝐶𝑉 2.77 × 10−9 × 200
𝐷𝑛 = 𝜌𝑆 = = = = 5.54 𝑛𝐶/𝑚2
𝐴 𝐴 1
Thus,
𝐷 5.54 × 10−9
𝐸1 = = = 1.25 × 104 𝑉/𝑚
𝜀𝑜 𝜀𝑟1 8.854 × 10−12 × 5

𝐷 5.54 × 10−9
𝐸2 = = = 6.25 × 104 𝑉/𝑚
𝜀𝑜 𝜀𝑟2 8.854 × 10−12 × 1

From which
𝑉1 = 𝐸1 𝑑1 = 1.25 × 104 × 10−3 = 125 𝑉
𝑉2 = 𝐸2 𝑑2 = 6.25 × 104 × 3 × 10−3 = 187.5 𝑉
Magnetic Field
The source of the steady magnetic field may be a permanent magnet, an electric field changing linearly
with time, or a dissect current i.e., a static magnetic field can originate from either a constant current
or a permanent magnet.
The electric field causes a force to be exerted on a charge which may be either stationary or in motion,
therefore, the steady magnetic field is capable of exerting a force only on a moving charge. The magnetic
field may be produced by moving charge. A magnetic field cannot arise from stationary charges and
cannot exert any force on a stationary charge.
Force on a Moving Charge
In an electric field, the definition of the electric field intensity shows as that force on a charged particle
as
𝐹̅ = 𝑄𝐸̅
The force is in the same direction as the electric field intensity and is directly proportional to both 𝐸̅ and
𝑄. A charged particle in motion in a magnetic field of flux density 𝐵̅ is found experimentally to
experience a force whose magnitude is proportional to the product of the magnet of the charge 𝑄, its
velocity 𝑣̅ and the flux density 𝐵̅ and to the sine of the angle between the velocity 𝑣̅ and 𝐵̅.
The direction of the force may be expressed as
𝐹̅ = 𝑄𝑣̅ × 𝐵̅
Therefore, the force on the moving particle arising from combined electric and magnetic fields is
obtained easily by superposition
𝐹̅ = 𝑄(𝐸̅ + 𝑣̅ × 𝐵̅)
This equation is known as the Lorentz force equation, and its solution required in determining electron
orbits in the magnetron, proton paths in the cyclotron, plasma characteristics in a
magnetohydrodynamic (MHD) generator or in general, charged particle motion in combined electric
and magnetic fields.
Example: The point charge 𝑄 = 18 𝑛𝐶 has a velocity of 5 × 106 𝑚/𝑠 in the direction 𝑎̅𝑣 = 0.60 𝑎̅𝑥 +
0.75 𝑎̅𝑦 + 0.30 𝑎̅𝑧 . Calculate the magnitude of the force exerted on the charge by the field

a. 𝐵̅ = −3𝑎̅𝑥 + 4𝑎̅𝑦 + 6𝑎̅𝑧 𝑚𝑇


b. 𝐸̅ = −3𝑎̅𝑥 + 4𝑎̅𝑦 + 6𝑎̅𝑧 𝑘𝑉/𝑚
c. 𝐵̅ and 𝐸̅ acting together
Solution
a. Using 𝐹̅ = 𝑄𝑣̅ × 𝐵̅
𝑣̅ = 𝑣 ∙ 𝑎̅𝑣 = 5 × 106 (0.60 𝑎̅𝑥 + 0.75 𝑎̅𝑦 + 0.30 𝑎̅𝑧 )

𝑣̅ = (3 𝑎̅𝑥 + 3.75 𝑎̅𝑦 + 6 𝑎̅𝑧 ) × 106

𝐹̅ = 18 × 10−3 (3 𝑎̅𝑥 + 3.75 𝑎̅𝑦 + 6 𝑎̅𝑧 ) × (−3𝑎̅𝑥 + 4𝑎̅𝑦 + 6𝑎̅𝑧 ) × 10−3

𝐹̅ = (54 𝑎̅𝑥 + 67.5 𝑎̅𝑦 + 108 𝑎̅𝑧 ) × 10−3 × (−3𝑎̅𝑥 + 4𝑎̅𝑦 + 6𝑎̅𝑧 ) × 10−3

𝑎̅𝑥 𝑎̅𝑦 𝑎̅𝑧


𝐹̅ = | 54 67.5 108| × 10−3
−3 4 6

𝐹̅ = (((67.5 × 6) − (108 × 4))𝑎̅𝑥 − ((54 × 6) − (108 × (−3))) 𝑎̅𝑦

+ ((54 × 4) − (67.5 × (−3))) 𝑎̅𝑧 ) × 10−3


𝐹̅ = ((405 − 432)𝑎̅𝑥 − (324 − 324)𝑎̅𝑦 + (216 − 202.5)𝑎̅𝑧 ) × 10−3

𝐹̅ = −27𝑎̅𝑥 + 13.5𝑎̅𝑧 𝑚𝑁

b. Using 𝐹̅ = 𝑄𝐸̅
𝐹̅ = 18 × 10−9 (−3𝑎̅𝑥 + 4𝑎̅𝑦 + 6𝑎̅𝑧 ) × 103

𝐹̅ = (54 𝑎̅𝑥 + 67.5 𝑎̅𝑦 + 108 𝑎̅𝑧 ) × 10−6

𝐹̅ = 54 𝑎̅𝑥 + 67.5 𝑎̅𝑦 + 108 𝑎̅𝑧 𝜇𝑁

c. Using 𝐹̅ = 𝑄(𝐸̅ + 𝑣̅ × 𝐵̅)


𝑎̅𝑥 𝑎̅𝑦 𝑎̅𝑧
𝑣̅ × 𝐵̅ = | 3 × 106 3.75 × 106 6 × 106 |
−3 × 10−3 4 × 10−3 6 × 10−3

𝑣̅ × 𝐵̅ = ((3.75 × 106 × 6 × 10−3 ) − (6 × 106 × 4 × 10−3 ))𝑎̅𝑥


− ((3 × 106 × 6 × 10−3 ) − (6 × 106 × (−3 × 10−3 ))) 𝑎̅𝑦
+ ((3 × 106 × 4 × 10−3 ) − (3.75 × 106 × (−3 × 10−3 ))) 𝑎̅𝑧

𝑣̅ × 𝐵̅ = ((22.5 − 24)𝑎̅𝑥 − (18 + 18)𝑎̅𝑦 + (12 + 11.25)𝑎̅𝑧 ) × 103

𝑣̅ × 𝐵̅ = (−1.5𝑎̅𝑥 − 32𝑎̅𝑦 + 23.25𝑎̅𝑧 ) × 103

𝐸̅ + 𝑣̅ × 𝐵̅ = (−3𝑎̅𝑥 + 4𝑎̅𝑦 + 6𝑎̅𝑧 ) × 103 + (−1.5𝑎̅𝑥 − 32𝑎̅𝑦 + 23.25𝑎̅𝑧 ) × 103

𝐸̅ + 𝑣̅ × 𝐵̅ = (−4.5𝑎̅𝑥 − 28𝑎̅𝑦 + 29.25𝑎̅𝑧 ) × 103

𝐹̅ = 18 × 10−9 (−4.5𝑎̅𝑥 − 28𝑎̅𝑦 + 29.25𝑎̅𝑧 ) × 103

𝐹̅ = (−81𝑎̅𝑥 − 504𝑎̅𝑦 + 526.5𝑎̅𝑧 ) × 10−6

𝐹̅ = −81𝑎̅𝑥 − 504𝑎̅𝑦 + 526.5𝑎̅𝑧 𝜇𝑁

Biot-Savart Law
This law states that at any point 𝑃, the magnitude of the magnetic field intensity produced by the
differential element is proportional to the product of the current, the magnitude of the differential
length, and the sine of the angle lying between the filament and a line connecting the filament to the
point 𝑃 at which the field is desired.

Also, the magnitude of the magnetic field intensity is inversely proportional to the square of the distance
from the differential element to the point 𝑃.
Mathematically,
𝐼 𝑑𝐿̅ × 𝑎̅𝑅 𝐼 𝑑𝐿̅ × 𝑅̅
̅=
𝑑𝐻 =
4𝜋𝑅 2 4𝜋𝑅 3
̅ are evidently amperes per metre (𝐴/𝑚). All elements making
The units of the magnetic field intensity 𝐻
up the complete current filament contribute to 𝐻 ̅ and must be included. If we locate the current element
at point 1 and describe the point P at which the field is to be determined as point 2, then
𝐼1 𝑑𝐿̅1 × 𝑎̅𝑅12
̅2 =
𝑑𝐻 2
4𝜋𝑅12
The summation leads to the integral form of the Biot-Savart law
𝐼 𝑑𝐿̅ × 𝑎̅𝑅 𝐼 𝑑𝐿̅ × 𝑅̅
̅=∮
𝐻 = ∮
4𝜋𝑅 2 4𝜋𝑅 3

Magnetic Field Strength of an Infinity Long Straight Conductor


To illustrate the application of Biot-savart law, consider an infinitely, long, straight, filamentary current
𝐼 along the z-axis in cylindrical coordinate as shown below.

Referring to the above figure, we should recognize the symmetry of this field, no variation with 𝑧 or with
𝜙 can exist. Point 2, at which we shall determine the field is therefore, chosen in the 𝑧 = 0 plane. The
field point 𝑟̅ , is therefore, 𝑟̅ = 𝜌𝑎̅𝜌 . The source point 𝑟̅ ′ is given by 𝑟̅ ′ = 𝑧 ′ 𝑎̅𝑧 and therefore,

𝑅̅12 = 𝑟̅ − 𝑟̅ ′ = 𝜌𝑎̅𝜌 − 𝑧 ′ 𝑎̅𝑧

So that
𝜌𝑎̅𝜌 − 𝑧 ′ 𝑎̅𝑧
𝑎̅𝑅12 =
√𝜌2 + 𝑧 ′ 2
We take 𝑑𝐿̅ = 𝑑𝑧 ′ 𝑎̅𝑧 and the Biot-savart law for differential length becomes

𝐼 𝑑𝑧 ′ 𝑎̅𝑧 × (𝜌𝑎̅𝜌 − 𝑧 ′ 𝑎̅𝑧 )


̅=
𝑑𝐻 3⁄
4𝜋(𝜌2 + 𝑧 ′ 2 ) 2

Since the current is directed towards increasing 𝑧 ′ , the limit the −∞ 𝑎𝑛𝑑 ∞ on the integral, and we have

𝐼 𝑑𝑧 ′ 𝑎̅𝑧 × (𝜌𝑎̅𝜌 − 𝑧 ′ 𝑎̅𝑧 )
̅= ∫
𝐻 3⁄
−∞
4𝜋(𝜌2 + 𝑧 ′ 2 ) 2


𝐼 𝑑𝑧 ′ 𝑎̅𝜙
̅=
𝐻 ∫
4𝜋 3
−∞
(𝜌2 + 𝑧 ′ 2 ) ⁄2
Since the integration is with respect to 𝑧 ′ , and 𝑎̅𝑧 is a constant and may be removed from under the
integral sign

𝐼 𝑟𝑎̅𝜙 𝑑𝑧 ′
̅=
𝐻 ∫
4𝜋 3
−∞
(𝜌2 + 𝑧 ′ 2 ) ⁄2


𝐼 𝑟𝑎̅𝜙 𝑧′
̅=
𝐻 [ ]
4𝜋 𝜌2 √𝜌2 + 𝑧 ′ 2
−∞

𝐼
̅=
𝐻 𝑎̅
2𝜋𝜌 𝜙
The magnitude of the field is not a function of 𝜙 or 𝑧, and it varies inversely with the distance from the
filament. The direction of the magnetic field intensity vector is circumferential. The streamlines are
therefore, circles about the filament and the field may be mapped in cross section as shown below.

The streamline of the magnetic field intensity about an infinitely long straight filament carrying a direct current 𝐼.

The separation of the streamline is proportional to the radius or inversely proportional to the magnitude
̅.
of 𝐻

Ampere’s Circuital Law


We have seen the Gauss’s law can be used to solve the same problem that Coulombs law can solve
whenever a high degree of symmetry is present. An analogous procedure exists in magnetic fields.
Here, the law that helps to solve the problem more easily is Ampere’s circuital law.
Ampere’s circuital law states that the line integral of the tangential component of the magnetic field
strength around a closed path is equal to the current enclosed by the path.

̅ ∙ 𝑑𝐿̅ = 𝐼𝑒𝑛𝑐𝑙𝑜𝑠𝑒𝑑
∮𝐻

At first glance, one would think that the law is used to determine the current 𝐼 by an integration. Instead,
the current is usually known and the law provides us the method of finding 𝐻 ̅.
̅ , there must be a considerable degree of symmetry in
In order to utilize Ampere’s law to determine 𝐻
the problem. Two conditions must be met:
i. ̅ is either tangential or normal to the path.
At each point of the closed path, 𝐻
ii. 𝐻 has the same value at all points of the path where 𝐻 ̅ is tangential.
̅ at radius 1 𝑚 from a long linear conductor is 1 𝐴𝑚−1. Find the current in
Example: The magnitude of 𝐻
the wire.
Solution
̅ ∙ 𝑑𝐿̅ = 𝐼
∮𝐻

𝐻 × 2𝜋𝑟 = 1 × 2𝜋 × 1 = 2𝜋
Curl
Let now apply Ampere’s circuital law to the perimeter of a differential surface element and drawn the
third and last of the special derivatives of vector analysis, the curl.
We shall choose rectangular coordinates, and an incremental closed path of sides ∆𝑥 and ∆𝑦 is selected
as shown in the figure below

An incremental closed path in rectangular coordinate is selected for the application of Ampere’s circuital law to
determine the spatial rate of change of 𝐻.
̅ at the centre of the small
We assume that some current, as yet unspecified, produce a reference value for 𝐻
rectangle
̅0 = 𝐻𝑥 𝑎̅𝑥 + 𝐻𝑦 𝑎̅𝑦 + 𝐻𝑧0 𝑎̅𝑧
𝐻

The closed line integral of 𝐻̅ about this path is then the sum of the four values of 𝐻
̅ ∙ ∆𝐿̅ on each side.
We choose the direction of traverse as 1 − 2 − 3 − 4 − 1, which corresponds to a current in the 𝑎̅𝑧
direction, and the first contribution is therefore,
̅ ∙ ∆𝐿̅)1−2 = 𝐻𝑦 1−2 ∆𝑦
(𝐻

The value of 𝐻𝑦 on this section of the path may be given in terms of the reference value 𝐻𝑦0 at the
centre of the rectangle, the rate of change of 𝐻𝑦 with 𝑥, and the distance ∆𝑥⁄2 from the centre to the
midpoint of side 1 − 2
𝛿𝐻𝑦 1
𝐻𝑦 1−2 = 𝐻𝑦0 + ( Δ𝑥)
𝛿𝑥 2
thus,
1 𝛿𝐻𝑦
̅ ∙ ∆𝐿̅)1−2 = (𝐻𝑦0 +
(𝐻 Δ𝑥) ∆𝑦
2 𝛿𝑥
along the next section of the path, we have
̅ ∙ ∆𝐿̅)2−3 = 𝐻𝑥 2−3 (−∆𝑥)
(𝐻

1 𝛿𝐻𝑥
̅ ∙ ∆𝐿̅)2−3 = − (𝐻𝑥0 +
(𝐻 Δ𝑦) ∆𝑥
2 𝛿𝑦
continuing for the remaining two segments and adding the results,
𝛿𝐻𝑦 𝛿𝐻𝑥
̅ ∙ 𝑑𝐿̅ = (
∮𝐻 − ) ∆𝑥∆𝑦
𝛿𝑥 𝛿𝑦
̅ in the direction 𝑎̅𝑛 is defined as
since the component of the curl of 𝐻
̅ ∙ 𝑑𝐿̅
∮𝐻
̅ ) ∙ 𝑎̅𝑛 = lim
(𝑐𝑢𝑟𝑙 𝐻
∆𝑆→0 ∆𝑆
In the coordinate systems, 𝑐𝑢𝑟𝑙 𝐻 ̅ is completely specified by its components along the three-unit
vectors. The components in the 𝑧 -direction is given by

∮𝐻̅ ∙ 𝑑𝐿̅
̅ ) ∙ 𝑎̅𝑧 =
(𝑐𝑢𝑟𝑙 𝐻 lim
∆𝑥→0,∆𝑦→0 ∆𝑥∆𝑦
therefore, from

𝛿𝐻𝑦 𝛿𝐻𝑥
̅ ∙ 𝑑𝐿̅ = (
∮𝐻 − ) ∆𝑥∆𝑦
𝛿𝑥 𝛿𝑦
this can be rearranged to give

∮𝐻̅ ∙ 𝑑𝐿̅ 𝛿𝐻𝑦 𝛿𝐻𝑥


lim = −
∆𝑥→0,∆𝑦→0 ∆𝑥∆𝑦 𝛿𝑥 𝛿𝑦
for the x, y components in the x, y – directions we have

∮𝐻̅ ∙ 𝑑𝐿̅ 𝛿𝐻𝑧 𝛿𝐻𝑦


lim = −
∆𝑦→0,∆𝑧→0 ∆𝑦∆𝑧 𝛿𝑦 𝛿𝑧
and

∮𝐻̅ ∙ 𝑑𝐿̅ 𝛿𝐻𝑥 𝛿𝐻𝑧


lim = −
∆𝑧→0,∆𝑥→0 ∆𝑧∆𝑥 𝛿𝑧 𝛿𝑥
̅ gives
combining the three components i.e., x, y and z components of the 𝑐𝑢𝑟𝑙 𝐻
𝛿𝐻𝑧 𝛿𝐻𝑦 𝛿𝐻𝑥 𝛿𝐻𝑧 𝛿𝐻𝑦 𝛿𝐻𝑥
̅=(
𝑐𝑢𝑟𝑙 𝐻 − ) 𝑎̅𝑥 + ( − ) 𝑎̅𝑦 + ( − ) 𝑎̅𝑧
𝛿𝑦 𝛿𝑧 𝛿𝑧 𝛿𝑥 𝛿𝑥 𝛿𝑦
̅
A third-order determinant can be written, the expansion of which gives the cartesian curl of 𝐻
𝑎̅𝑥 𝑎̅𝑦 𝑎̅𝑧
𝛿 𝛿 𝛿
̅ = ||
𝑐𝑢𝑟𝑙 𝐻 ||
𝛿𝑥 𝛿𝑦 𝛿𝑧
𝐻𝑥 𝐻𝑦 𝐻𝑧
̅ can
The elements of the second row are the components of the 𝑑𝑒𝑙 operator. This suggests that ∇ × 𝐻
̅
be written for 𝑐𝑢𝑟𝑙 𝐻 .
̅ in cylindrical and spherical coordinate can be derived in the same manner as
Expression for the 𝑐𝑢𝑟𝑙 𝐻
above, though with more difficulty as

1 𝛿𝐻𝑧 𝛿𝐻𝜙 𝛿𝐻𝑟 𝛿𝐻𝑧 1 𝛿(𝑟𝐻𝜙 ) 𝛿𝐻𝑟


̅=(
𝑐𝑢𝑟𝑙 𝐻 − ) 𝑎̅𝑟 + ( − ) 𝑎̅𝜙 + ( − ) 𝑎̅𝑧
𝑟 𝛿𝜙 𝛿𝑧 𝛿𝑧 𝛿𝑟 𝑟 𝛿𝑟 𝛿𝜙
And

1 𝛿(𝐻𝜙 sin 𝜃) 𝛿𝐻𝜃 1 1 𝛿𝐻𝑟 𝛿(𝑟𝐻𝜙 ) 1 𝛿(𝑟𝐻𝜙 ) 𝛿𝐻𝑟


̅=
𝑐𝑢𝑟𝑙 𝐻 ( − ) 𝑎̅𝑟 + ( − ) 𝑎̅𝜃 + ( − ) 𝑎̅𝜙
𝑟 sin 𝜃 𝛿𝜃 𝛿𝑧 𝑟 sin 𝜃 𝛿𝜙 𝛿𝑟 𝑟 𝛿𝑟 𝛿𝜃
, respectively.

Frequently useful are two properties of the curl operator.

i. The divergence of a curl is the zero scalar, i.e., ∇ ∙ (∇ × 𝐻 ̅.


̅ ) = 0 for any vector field 𝐻
ii. The curl of a gradient is the zero vector, i.e., ∇ × (∇𝑓) = 0 under static condition, 𝐸 ̅ = −∇𝑉,
and so ∇ × 𝐸̅ = 0.
Relationship of 𝑱̅ and 𝑯
̅
̅ )𝑥 may be rewritten as
In view of Ampere’s circuital law, the defining equation for (∇ × 𝐻

∮𝐻̅ ∙ 𝑑𝐿̅
̅ )𝑎̅𝑥 =
(∇ × 𝐻 lim
∆𝑦→0,∆𝑧→0 ∆𝑦∆𝑧
𝐼𝑥
̅ )𝑎̅𝑥 =
(∇ × 𝐻 lim = 𝐽𝑥
∆𝑦→0,∆𝑧→0 ∆𝑦∆𝑧
𝑑𝐽
̅ and the
Where 𝐽𝑥 = 𝑑𝑆𝑥 is the area density of 𝑥 directed current. Thus, the 𝑥 components of 𝑐𝑢𝑟𝑙 𝐻
current density 𝐽 ̅ are equal at any point.
Similarly, for 𝑦 and 𝑧 components, so that
̅ = 𝐽̅
∇×𝐻
This is one of Maxwell’s equation for static fields.
Example: Calculate the value of the vector current density in
i. rectangular coordinate at 𝑃(2,3,4), if 𝐻 ̅ = 𝑥 2 𝑧𝑎̅𝑦 − 𝑦 2 𝑥𝑎̅𝑧 .
ii. ̅ = 2 (cos 0.2)𝑎̅𝜌 .
cylindrical coordinate at 𝑃(1.5, 90° , 0.5), if 𝐻
𝜌
1
iii. ̅=
spherical coordinate at 𝑃(2, 30° , 20° ), if 𝐻 𝑎̅
sin 𝜃 𝜃

Solution
i. rectangular coordinate
𝑎̅𝑥 𝑎̅𝑦 𝑎̅𝑧
𝛿 𝛿 𝛿
̅ = 𝐽 ̅ = ||
∇×𝐻 ||
𝛿𝑥 𝛿𝑦 𝛿𝑧
𝐻𝑥 𝐻𝑦 𝐻𝑧

𝑎̅𝑥 𝑎̅𝑦 𝑎̅𝑧


𝛿 𝛿 𝛿
𝐽 ̅ = || |
𝛿𝑥 𝛿𝑦 𝛿𝑧 |
0 𝑥2𝑧 −𝑦 2 𝑥

𝛿 (−𝑦 2 𝑥) 𝛿 (𝑥 2 𝑧) 𝛿 (−𝑦 2 𝑥) 𝛿 (0) 𝛿 (𝑥 2 𝑧) 𝛿 (0)


𝐽̅ = ( − ) 𝑎̅𝑥 − ( − ) 𝑎̅𝑦 + ( − ) 𝑎̅𝑧
𝛿𝑦 𝛿𝑧 𝛿𝑥 𝛿𝑧 𝛿𝑥 𝛿𝑦

𝐽 ̅ = (−2𝑦𝑥 − 𝑥 2 )𝑎̅𝑥 − (−𝑦 2 − 0)𝑎̅𝑦 + (2𝑥 − 0)𝑎̅𝑧

𝐽 ̅ = −(2𝑦𝑥 + 𝑥 2 )𝑎̅𝑥 + 𝑦 2 𝑎̅𝑦 + 2𝑥𝑎̅𝑧

At 𝑃(2,3,4),

𝐽 ̅ = −(2(3)(2) + 22 )𝑎̅𝑥 + 3𝑎̅𝑦 + 2(2)𝑎̅𝑧

𝐽 ̅ = −16𝑎̅𝑥 + 3𝑎̅𝑦 + 4𝑎̅𝑧 𝐴

Magnetic Flux Density (𝑩 ̅)


̅ ̅
Like 𝐷, the magnetic field strength 𝐻 depends only on moving charges and is independent of the
̅ is the magnetic flux density 𝐵̅, which is given by
medium. The force field associated with 𝐻
𝐵̅ = 𝜇𝐻
̅

Where 𝜇 = 𝜇𝑜 𝜇𝑟 is the permeability of the medium. The unit of 𝐵̅ is the Tesla.


The free space permeability 𝜇𝑜 has a numerical value of 4𝜋 × 10−7 𝐻/𝑚.
Magnetic flux Φ through a surface is defined as

Φ = ∫ 𝐵̅ ∙ 𝑑𝑆̅
𝑐

The unit of magnetic flux is the Weber (Wb). The magnetic flux lines are closed and do not terminate
on a magnetic charge. For this reason, Gauss law for the magnetic field is

∫ 𝐵̅ ∙ 𝑑𝑆̅ = 0
𝑐

And the application of the divergence theorem shows that


∇ ∙ 𝐵̅ = 0
This is another Maxwell’s equation. Collecting these equations in differential form, we than have for
static electric field and steady magnetic fields
̅ = 𝜌𝑣
∇∙𝐷
∇ ∙ 𝐸̅ = 0
̅ = 𝐽̅
∇∙𝐻
∇ ∙ 𝐵̅ = 0
̅ to 𝐸̅ and 𝐵̅ to 𝐻
To these equations, we may add the two expressions relating 𝐷 ̅ in free space,
̅ = 𝜀𝑜 𝐸̅
𝐷
𝐵̅ = 𝜇𝑜 𝐻
̅

In integral form, we have the corresponding set of four integral equations

̅ ∙ 𝑑𝑆 = ∫ 𝜌𝑣 𝑑𝑣
∮𝐷
𝑣𝑜𝑙

∮ 𝐸̅ ∙ 𝑑𝐿̅ = 0

̅ ∙ 𝑑𝐿̅ = 𝐼 = ∫ 𝐽 ̅ ∙ 𝑑𝑆
∮𝐻
𝑐

∮ 𝐵̅ ∙ 𝑑𝑆 = 0
𝑆

You might also like