0% found this document useful (0 votes)
6 views

Lecture Notes 5-8 (1)

engineering

Uploaded by

benten10prince
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
6 views

Lecture Notes 5-8 (1)

engineering

Uploaded by

benten10prince
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 24

2 Electric Field

2.1. Importance of Electrostatics


Electrostatics is a fascinating subject that has grown up in diverse areas of application. Electric
power transmission, X-ray machines, and lightning protection are associated with strong electric
fields and will require a knowledge of electrostatics to understand and design suitable equipment.
The devices used in solid-state electronics are based on electrostatics. These include resistors,
capacitors, and active devices such as bipolar and field effect transistors, which are based on
control of electron motion by electrostatic fields.
Almost all computer peripheral devices, with the exception of magnetic memory, are based on
electrostatic fields. Touch pads, capacitance keyboards, cathode-ray tubes, liquid crystal displays,
and electrostatic printers are typical examples.
In medical work, diagnosis is often carried out with the aid of electrostatics, as incorporated in
electrocardiograms, electroencephalograms, and other recordings of organs with electrical activity
including eyes, ears, and stomachs.
In industry, electrostatics is applied in a variety of forms such as paint spraying, electro-deposition,
electrochemical machining, and separation of fine particles. Electrostatics is used in agriculture to
sort seeds, direct sprays to plants, measure the moisture content of crops, spin cotton, and speed
baking of bread and smoking of meat.

2.2. Coulomb’s law


Coulomb’s law states that, the force between two small charges Q1, and Q2 (idealized as point
charges of zero size) is proportional to their magnitudes and inversely proportional to the square
of the distance R12 between them as shown in figure 2.1,

Figure 2.1 Coulombs law illustration

The force acts along the line joining the charges in the same or opposite direction of the unit vector
aR12 and is attractive if the-charges are of opposite sign and repulsive if like charged. The force
F12 on charge Q2 due to charge Q1 is equal in magnitude but opposite in direction to the force F21,
on Q1, and the net force on the pair of charges being zero.
1 Q1Q2 
F12   F21  aR12 (2.1)
4 0 R 2
Where
F is the force between the charges (N), q1 and q2 are the magnitudes of the two charges (C), ε0 is
a material constant (F m−1), r12 is the distance between the charges (m), and r̂ is a unit vector
acting in the direction of the line joining the two charges.

(2.2)

(2.3)

Substituting for aR12 in equation 2.1


1 Q1Q2  1 Q1Q2 (r2  r 1 )
F12   F21  R12 or F12   F21  3
(2.4)
4 0 R 3 4 0 r2  r 1

If the charge Q1 exists alone, it feels no force. If we now bring charge Q2 within the vicinity of Q1,
then Q2 feels a force that varies in magnitude and direction as it is moved about in space and is
thus a way of mapping out the vector force field due to Q1. A charge other than Q2 would feel a
different force from Q2 proportional to its own magnitude and sign.
Example 17) Determine the force between two identical charges, of magnitude 10 pC, separated by
a distance of 1mm situated in free-space. What is the force if the separation is reduced
to 1μm?
1 Q1Q2  10 1012 10  1012  
F 2
a12  2
a12  0.9 106 a12 N
4 R 4    8.854 1012  1103 

If we reduce the separation to 1μm, the force increases to 0.9N.


Example 18) Point charges 1 mC and - 2 mC are located at (3, 2, -1) and (-1, -1,4), respectively.
Calculate the electric force on a 10 nC charge located at (0, 3, 1) at that point.
Solution
Q1= 1 mC=1x10-3C, Q2=-2 mC=-2x10-3C and Q=10 nC=10x10-9C
− = (0,3,1) − (3,2, −1) = −3 + +2

| − |= (−3) + 1 + 2 = 3.742
− = (0,3,1) − (−1, −1,4) = +4 −3

| − |= (1) + 4 + (−3) = 5.099


( − ) ( − )
= +
4 | − | 4 | − |
10 × 10 × 1 × 10 (−3 + +2 ) 10 × 10 × (−2) × 10 ( + 4 −3 )
= +
10 10
4 |3.742| 4 |5.099|
36 36

0.09(−3 + +2 ) 0.18(( +4 − 3 ))
= 3 − 3
|3.742| |5.099|
= −0.00515 + 0.00172 + 0.00343 − (0.00136 + 0.00543 − 0.00407 )

=− . − . + .

2.3. The Electric Field Intensity or Electric Field Strength


The electric field intensity (or electric field strength) E is the force per unit charge when placed in
the electric field.
Considering Q2 as the test charge, the electric field intensity due to Q1 at the position of Q2 is
defined as;
 F Q1 
E2  lim 2  R12 volts/m (2.5)
Q2 0 Q
2 4 0 R 2
In general, from equation 2.4
 1 Q1 (r2  r 1 )
E 3
volts/m (2.6)
4 0 r2  r 1

Consider a test charge QP, is placed at point P, in Figure 2.3. In the vicinity of N charges, the test
charge will feel a force
 
FP  qP EP (2.7)

Figure 2.2 The electric field due to a collection of point charges

Ep is the vector sum of the electric fields due to all the N-point charges,

(2.8)

Example 19) Determine the electric field strength at a distance of 0.5m from an isolated point
charge of +10 μC. If an identical charge is placed at this point, determine the force it
experiences. Assume that the charge is in air.
Example 20) Two-point charges are a distance a apart along the z axis as shown in Figure 2-8. Find
the electric field at any point in the z =0 plane when the charges are:
i. both equal to q
ii. of opposite polarity but equal magnitude ±q. This configuration is called an
electric dipole.
Example 21) Determine electric field intensity for the particle in example 18
Solution
= −0.00651 − 0.00371 + 0.0075
−0.00651 − 0.00371 + 0.0075
= =
10 × 10
=− − + /

Assignment 1
A practical application of electrostatics is in electrostatic separation of solids. For example, Florida
phosphate ore, consisting of small particles of quartz and phosphate rock, can be separated into its
components by applying a uniform electric field as in Fig. Asg. 1. Assuming zero initial velocity
and displacement, determine the separation between the particles after falling 80 cm. Take E = 500
kV/m and Q/m = 9 µC/kg for both positively and negatively charged particles.

Figure. Asg. 1
Example 22) 1. Determine the flux radiating from a positive point charge of magnitude 100 pC.
2. What is the flux density at a distance of 10mm from the charge?
3. Determine the flux that flows through an area of 200mm2 on the surface of a 1-m
radius Gaussian sphere.
4. Repeat (3) if a negative charge of the same magnitude replaces the positive charge.

2.4. Charge Distribution


In typical situations, one coulomb of total charge may be present requiring 6.25 x 1018 elementary
charges ( e  1.60  1019 coul ). When dealing with such a large number of particles, the discrete
nature of the charges is often not important and we can consider them as a continuum. We can then
describe the charge distribution by its density.
2.4.1. Line, Surface, and Volume Charge Distributions
Charges can distribute themselves on a line with line charge density ρL (C/m), on a surface with
surface charge density ρs (C/m2) or throughout a volume with volume charge density ρv (C/m3).
Consider a distribution of free charge dq of differential size within a macroscopic distribution of
line, surface, or volume charge as shown in Figure 2.4.
Figure 2.3 Charge distributions. (a) Point charge; (b) Line charge; (c) Surface charge; (d) Volume charge.

The total charge Q within each distribution is obtained by summing up all the differential elements.
This requires an integration over the line, surface, or volume occupied by the charge.

The electric field intensity due to each of the charge distributions pL, ps, and pv may be regarded
as the summation of the field contributed by the numerous point charges making up the charge
distribution. Thus by replacing Q in eq. (2.5) with charge element dQ = pL dl, ps dS, or pv dv and
integrating, we get

Example 23)
Example 24)
2.5. Electric Flux and Electric Flux Density
One definition of flux is that it is the flow of material from one place to another. If we adapt this
to electrostatics, we can say that a positive charge is a source of electric flux, and a negative charge
acts as a sink.
Consider a point charge, Q1 surrounded an imaginary
sphere as shown in figure 2.3.
The flux density, D may be defined as electronic charge
flux per unit area. Gauss’ law states that the flux
through any closed surface is equal to the charge
enclosed by that surface.
ψ=
That is;
Figure 2.4 Flux density ,
ψ= dψ = ∙

From volume distribution charge,


,
=

Therefore;
, ,
ψ= ∙ =

By applying divergence theorem


, ,
∙ = ∇∙

Hence
, ,
∇∙ = ⇒ ∇∙ =

Example 25)
2.6. Electric Potential
Consider figure 2.7, we have a positive test charge
of 1C at a distance d1 from the fixed negative
charge, −q1. This test charge will experience an
attractive force whose magnitude we can find
from Coulomb’s law. Now, if we move the test
charge from position 1 to position 2, we have to
do work against the field.
From Coulomb's law, the force on Q is F = QE so
that the work done in displacing the charge by dr
Figure 2.5 Potential energy in electrostatic field
is
 
dW   F  dr  QE  dr (2.9)
The negative sign indicates that the work is being done by an external agent. Thus the total work
done, or the potential energy required, in moving Q from 1 to 2 is
2

W  Q  E  dr (2.10)
1

Dividing W by Q in eq. (2.10) gives the potential energy per unit charge. This quantity, denoted
by V12, is known as the potential difference between points 1 and 2. Thus
W
2

V12     E  dr (2.11)
Q 1

We can move from position 1 to position 2 in very tiny steps so that the E field hardly varies with
each step. With each step we take, we will do a small amount of work against the field. To find
the total amount of work done, and hence the potential difference, we can integrate Equation (2.11)
with respect to r, with d1 and d2 as the limits. Thus,
d2 d2
Q1
V 12    Edr    4 r 2
dr
d1 d1
d2 d2
(2.12)
Q1 1 Q1  1  Q1  1 1 
 d1 r 2 dr   2 
     
4 r 2 4 r  r d 1  2
4 r  d1 d 2 

The this equation can be recast as


Q1 1 Q1 1
V1  V2   (2.13)
4 r d1 4 r 2 d 2
2

From which we can infer that;

Q1 1 Q1 1
V1  and V2  (2.14)
4 r 2 d1 4 r 2 d 2
Example 26) Determine the absolute potential at a distance of 0.2m from an isolated point charge
of 10 μC. Hence, determine the potential difference between this point and another at
10m from the charge.
Solution
The absolute potential is defined as the work done against the field in moving a positive 1C test
charge from initially to a point in the field. So, the small amount of work done, dV, in moving
distance dr is
dV   force  dr  1 E  dr   Edr
Thus, the total work done in moving the charge from infinity to 0.2m from the fixed charge, the
potential, is
v 0.2 0.2
10 106 10 106 1
0 dV    4 r 2 dr   

4 r
10 106 1
  4.5 105 V
4 0.2
By following a similar procedure, the potential at 10m from the charge is
V  9  103 V
Thus, the potential difference between 0.2 and 10m is
V12  4.41105 V

2.7. Poisson's and Laplace's Equations


Poisson's and Laplace's equations are easily derived from Gauss's law (for a linear material
medium)
Recall
= , ∇∙ = , and = −∇
∇∙ = ∇. =
Substituting for E

∇. (−∇ ) = ⇒∇ =−

This is known as Poisson's equation. A special case of this equation occurs when pv = 0, l.e
∇ = 0 which is known as ′
3 Magnetic Field
3.1 Magnetostatics
Similar to electrostatics, Coulomb found that the force between two magnetic poles decreases as
the inverse of the square of the distance separating them, i.e.,
 pp
F  1 22 rˆ (3.1)
kr
Where; F is the vector force between the two poles (N), p1 and p2 are the strengths of the magnetic
poles (Wb), k is a constant of proportionality, r is the distance between two poles (m) and r̂ is the
unit vector acting in the direction of the line joining the two charges.
The force is repulsive if the poles are alike and attractive if the poles are dissimilar see figure 3.1.

Figure 3.1 Two separate magnetic monopoles in free-space

We can extract a factor of 4π from the constant k to give


 pp
F  1 22 rˆ (3.2)
4 r
Where μ is the permeability – a material property.
The idea of magnetic field strength can be introduced in the same way that we introduced electric
field strength. The force on the pole p2 is
 
F  p2 H (3.3)
Where H is the magnetic field strength due to pole p1 given by

 p1
H 2
rˆ N W 1 (3.4)
4 r
Similarly, if we adapt Gauss’ law to magnetostatics, we can say that the magnetic flux emitted by
a pole is equal to the strength of the pole. Thus, we can define the magnetic flux density as
 p
B  1 2 r Wb m 2 . (3.5)
4 r
We can combine Equations (3.4) and (3.5) to give
 
B  H (3.6)
The permeability of free-space μ0 is with value 4π × 10−7 H m−1.
Example 27) A single 10 µWb magnetic monopole is situated in air. Calculate the magnetic field
strength at a distance of 0.5m from the monopole. In addition, find the flux density
and the force on an identical monopole at the same distance.
3.2 Magnetostactics conventions
o A cross denotes current flowing into the page. A dot denotes current flowing out of the
page.
o The right-hand corkscrew rule gives the direction of the flux.
o Clockwise rotation of the field denotes a south pole. Anticlockwise rotation of the field
gives a north pole.

3.3 The Biot-Savart Law


Consider the current-carrying conductor shown below. The conductor generates a magnetic field
that is coaxial to the wire. If we place an imaginary unit north pole at a point P, distance r from a
small elemental section of the wire of length dl, the wire will experience a force that will tend to
push it to the left. In addition, there will be an equal and opposite force on the North Pole due to
the field surrounding the wire.

The north pole (as a point source), emits magnetic flux in a radial direction given as
 p
BN  N 2 rˆ (3.7)
4 r
Direct experimental measurement shows that the force on a current-carrying conductor placed in
a magnetic field is given by
 
F  BIl (3.8)

Where B is the flux density of the magnetic field in which the wire is placed, I is the current
flowing through the wire and l is the length of the wire.
By combining Equations (3.7) and (3.8), we find that the force on the element dl due to the field
emitted by the north pole is
 p
dF  N 2 Idl (3.9)
4 r
Let us now turn our attention to the force on the north pole produced by the current element. The
current element formed by the current I and the length dl produces a magnetic field strength of
dH1, at the north pole. By applying Equations (3.3) and (3.4), the force on the north pole due to
the current element is

dF  pN dH1 (3.10)
By equating Equations (3.9) and (3.10), we get
pN   Idl
Idl  p N dH 1,  dH1 =  (3.11)
4 r 2 4 r 2
Where ϕ is the unit vector acting into the page. To find the total field produced by the wire, we
should integrate this equation with respect to the length.
However, Equation (3.11) only gives the field at a point directly opposite the current element we
are considering. To find the field we require, we
must do some resolving of components.
If we draw a line from the point P to the current
element, we find that it makes an angle to the
current element of θ.
So, if we modify Equation (3.11) by a factor sin θ,
we will get
Idl
dH  sin  (3.12)
4 r 2
Equation (3.12) is known as the Biot-Savart law. It is easy to notice that eq. (3.12) is better put in
vector form as
 
Idl  aR Idl  r
dH   (3.13)
4 r 2 4 r 3
Biot-Savart's law states that the magnetic field intensity dH produced at a point P by the
differential current element I dl is proportional to the product I dl and the sine of the angle (θ)
between the element and the line joining P to the element and is inversely proportional to the
square of the distance r between P and the element.

3.4 Current distribution


We can have different current distributions: line current, surface current, and volume current. If
we define K as the surface current density (in amperes/meter) and J as the volume current density
(in amperes/meter square), the source elements are related as
Idl  Kds  Jdv (3.14)
Thus in terms of the distributed current sources, the Biot-Savart law as in eq. (3.13) becomes

Example 28) Determine the field due to a straight current carrying filamentary conductor of finite
length AB as in Figure below. Assume that the conductor is along the z-axis with its
upper and lower ends respectively subtending angles α2 and α1 at P, the point at which
H is to be determined.
Solution
If we consider the contribution dH at P due to an element dl at (0, 0, z),

Idl  r 
dH  3
But dl  dza z and r = R =  a   za z
4 r

dl  R =  dza and r   2  z 2
I  dz
Hence H = a
2 
2 2
4
  z 
Letting z   cot  , dz    cos ec 2 d then,

Or

This expression is generally applicable for any straight filamentary conductor of finite length.
Notice from the equation that H is always along the unit vector aϕ (i.e., along concentric circular
paths) irrespective of the length of the wire or the point of interest P.
Example 29) A circular loop located on x2 + y2 = 9, z = 0 carries a direct current of 10 A along aϕ.
Determine H at (0, 0, 4) and (0, 0, -4).
3.5 Ampere's Circuit Law—Maxwell's Equation
Ampere's circuit law states that the line integral of the tangential component of H around a dosed
path is the same as the net current Ienc enclosed by the path.
In other words, the circulation of H equals Ienc that is,

 H  dl  I
 enc (3.15)

By applying Stoke's theorem to the left-hand side of eq. (3.15), we obtain

L
 H  dl ( H ).dS
I enc  
s
Also, I enc J  dS
s
(3.16)

Hence

3.6 Magnetic Flux Density—Maxwell's Equation


The magnetic flux density B is similar to the electric flux density D. As D = εE in free space, the
magnetic flux density B is related to the magnetic field intensity H according to
 
B  H (3.17)
The permeability of free-space μ0 is with value 4π × 10−7 H m−1.
The magnetic flux through a surface S is given by
 B  dS (3.18)
S

An isolated magnetic charge does not exist. Thus the total flux through a closed surface in a
magnetic field must be zero; that is,

B  dS  0 (3.19)
S

By applying the divergence theorem to eq. (3.8), we obtain

B  dS  B  0 (3.20)
S v

Or

(3.21)

3.7 Magnetic Scalar and Vector Potentials


Just as = −∇ , the magnetic scalar potential Vm (in amperes) as related to H according to is
Thus the magnetic scalar potential Vm is only defined in a region where J = 0 . We should also note
that Vm satisfies Laplace's equation just as V does for electrostatic fields; hence,

We can define the vector magnetic potential A (in Wb/m) such that

Example 30) Given the magnetic vector potential = Wb/m, calculate the total magnetic
flux crossing the surface ∅ = , ≤ ≤ , ≤ ≤ .
Solution
Homework

3.8 Electromagnetic Flux, Flux Density and Field Strength


For electromagnetic fields, we can adapt Gauss’ law to state that the magnetic flux generated by a
current element is equal to the product of the current and element length.
p  Idl Wb (3.22)
If we consider the isolated current element, I dl, we have a fractional magnetic field strength, dH,
of acting into the page.
Idl
dH  sin  A/m (3.23)
4 r 2
If we introduce a north pole at the point P, it will experience a force of
 Idl
dF  sin  pN (3.24)
4 r 2
The force on a magnetic pole is directly dependent on the magnetic field strength produced by the
wire. This agrees with our model of electrostatics. As B = μH, the fractional magnetic flux density
is given by
 Idl
dB   sin  (3.25)
4 r 2
Equation 3.16 is better put in vector form as
 Idl
dB   sin  (3.26)
4 r 2
We should note that if we introduce a current element of strength p, at the point P, the force on
this element will be given by
 Idl  
dF   sin  pi or dF  dBpi (3.27)
4 r 2
Thus, the force on a current element is directly dependent on the magnetic flux density produced
by the wire, and not the magnetic field strength.

You might also like