Download as pdf or txt
Download as pdf or txt
You are on page 1of 56

CHAPTER 2

MAGNETOSTATIC FIELD
 CHAPTER OBJECTIVES

1) Understand the qualitative features of the interactions of charges moving with constant
velocity.
2) Know and be able to determine the nature of the magnetic field for various current
configurations.
3) Be able to understand the force on a current carrying conductor in magnetic field.
4) Be able to describe the interaction of parallel current carrying conductors.
1) Describe and discuss the industrial use of magnetostatic fields.

Chapter Contains

Chapter 2
2.1 Introduction
2.2 Biot-Savart‘s Law
2.3 Ampere‘s Circuital Law
2.4 Gauss‘s Law for Magnetic Field
2.5 Magnetic Scalar and Vector Potentials
2.6 Calculation of Magnetic Flux
2.7 Lorentz Force
2.8 Magnetic force between current carrying conductors
2.9 Boundary Conditions for Magnetostatic Field
2.10 Energy in Magnetic Field
2.11 Calculation of Inductance
2.12 Application of Magnetostatic Field
Summary
Sample Questions
2.1 INTRODUCTION
Magnetic phenomena are also the basis of much of our modern technology. The
development of electric generators and motors, transformers, microphones, compasses,
telephone bell ringers, television focusing controls, advertising displays, magnetically
levitated high-speed vehicles, magnetic tapes to store visual and audio information,
magnetic separators, and so on, involve magnetic phenomena and play an important role
in our daily life.
In this chapter, we will consider the magnetostatic field associated with current carrying
conductor. Charges with constant velocity are the source of this magnetic field.
There are two major laws governing magnetostatic fields: (i) Biot-Savart‘s law, and (ii)
Ampere‘s circuital law. Biot-Savart‘s law is the general law of magnetostatics. Ampere‘s
law is the special case of Biot-Savart‘s law and is easily applied in problems involving
symmetrical current distribution.

2.2 BIOT-SAVART’S LAW


Biot and Savart quantified Ampere‘s measurements by showing that the magnetic field
qu sin 
intensity H at a distance R away from a charge q moving with velocity u is H
R2
where the proportionality constant in SI unit is 1/4. Therefore,
1 u  aR
H q (2.1)
4 R2
where H and u are the magnitude of H and u,  is the angle between u and R (here R is a
vector pointing from the charge to the point where the field is being found) as shown in
the following figure, and aR is the unit vector acting along R.

In fact, we are often interested not the field of moving charge, but for that of a current
carrying conducting wire. Biot and Savart proposed a mathematical formula for the
magnetic field due to a single current element (a small conducting wire of length dl
having current I through it) as follows:
The problem geometry is shown in a figure given bellow. Suppose cross-sectional area of
the wire is A and v is the density of charge in the wire. Assume these charges move with
velocity u. Then the charge crossing any cross-section A in one second=uAv=I. Idl=
uAdlv=uq.
Therefore the magnetic field dH at point P, as shown in the above figure, due a current I
through a differential length dl of a wire (usually called current element Idl) is given by
the following relation after substituting qu=Idl in equation (2.1):

1 I dl   a R 1 I dl   (r  r )
dH   (2.2)
4 R 2
4 r  r
3

The resultant field is found experimentally to be normal to the plane containing the
vectors dl and R. This expression (2.2) is known as the ‗Biot-Savart law‘. The unit of
magnetic field intensity H is ampere-turn/m.
By the use of a suitable integration, Biot-Savart law can be used to calculate the H-field
resulting from a current carrying conductor of any length L which can be decomposed
into an infinite number of back to back connected current elements:

I dl   (r  r )
H
4  r  r
3
(2.3)

The Biot-Savart law can also be expressed in terms of distributed current sources, shown
in the following figure, such as the surface current density K, measured in A/m, or the
volume current density J, measured in A/m2.

(a) (b)
The total current flowing across the surface of the conductor shown in figure (a) is
I   Kdl and the total current crossing the cross-section S of the cylinder shown in
l

figure (b) is I   J  d s .
S

Thus the differential current element Idl, where dl is in the direction of the current, may
be expressed in terms of the surface current density K or volume current density J as Idl
=Kds=Jdv. Therefore, the Biot-savart law in terms of surface current K and volume
current J can be obtained as:

1 K  (r  r)
H
4  r  r
3
ds (2.4a)

1 J  (r  r )
or H 
4  r  r
3
dv  (2.4b)

The direction of magnetic field can also be found by using Right Hand Thumb’s Rule
which states in the following way: Grasp the wire with right hand; if the thumb points in
the direction of current, the fingers will curl around the wire in the direction of H or vice
versa.
It is to be noted that the magnetic flax density B=H where  is the permeability of the
medium (the permeability of the free space 0=410-7 H/m). Thus equations (2.1~2.4)
can also hold for B.

Example 2.1: Determine B-field due to an infinitely long, straight current carrying wire.

Solution: We would like to determine B at P a perpendicular distance  from the wire as


shown in the following figure.
Consider dB produced by a current element of length dl=dz at (0, 0, z) and lies along
the z axis and a distance R from P as shown in the figure.

Assume the point P (x, y, 0).


a ρ  z a z
R  r  r   xa x  ya y  z a z  a ρ  z a z  R   2  z  2  
1/ 2
and a R 
 2
 z2 
1/ 2

μI dl  r  r  μI dz a z  a ρ  z a z  μI dz a φ
4  r  r  3 4  4   2  z  2 3 / 2
B  3/ 2

 2  z2
Let, z=tan. Then dz=sec2d and  2  z  2  =2sec2.
LU
μI dza φ μI z
B 
4  2
 z2
3/ 2

4  2  z2 1/ 2
aφ .
LL

where z=LU and z=LL are the upper and lower position of the wire along z-axis. For an
infinitely long wire the limits of the integral are LU= and LL= - . Hence
μI
 B aφ
2 
The resultant B-field, therefore, is of the form of concentric circles around the wire.
I
Example 2.2: Using Biot-Savart law, show that H  (sin  2  sin 1 )aˆ at point 2 in
4
the following figure.

Assume a current element Idl at a height z and point 2 is on the xy-plane at z=0. Then
r=a, r=zaz, r  r=(a  zaz), and | r  r|=  2  z 2  .
1/ 2

Now from Biot-Savat law, we obtain


I dl  r  r μI dz a z  a ρ  z a z  I dz a
4  r  r 3 4  
H  
 2  z2
3/ 2
4 2
 z2
3/ 2

Following the procedure of integration in the last problem, we obtain


LU
I z I
H a  (sin  2  sin 1 )a
4  2  z 
2 1/ 2 4
LL

where z=LU and z=LL are the upper and lower position of the wire along z-axis.
Example 2.3: Derive an expression for the magnetic field at a point P(x, y) on the x-y
plane due to two long straight parallel wires directed along z-axis and placed at y=d/2 and
y= -d/2. Assume the currents in the wires are opposite to each other.

Solution:
Let us first consider the field due to a single infinitely long conductor which is placed at
y=-d/2 as shown in the following figure:
z

I
dz r1={x2+ (y+d/2)2+ z2}1/2={12+ z2}1/2
-I
z r1
-y y
1
P(x,y)

x
I
Using the result for infinitely long wire we have B  a  . It can be expressed in
2 
rectangular coordinate system as B  B x a x  B y a y  B z a z where
I I I I I
Bx  a  a x   sin  , B y  a  a y  cos  and B z  a  a z  0
2 2 2 2 2
For this particular problem, the field due to the left conducting wire is
I y  d / 2 I x
Bx   , By  and B z  0
21 1 21 1
And the field due to the right conducting wire is
I d / 2  y I x
Bx   , By   and Bz  0
22  2 22  2
Along the line midway between the two wires (x=y=0 and 1=2=d/2),
2I
Bx   , B  0 and Bz  0.
d y

Example 2.4: Assume that the wire in the following figure has a current I through it in
the direction shown. What is the magnitude of the magnetic field produced at point C due
to the straight segments, (b) the semicircular arc and (c) the entire wire?
y

x
Solution:
a. Looking at the figure, we see that the straight sections have infinitesimal lengths
that lie parallel or antiparallel to R, the vector that goes from dl to point C. Therefore,
dl×aR = 0 in both cases for every dl along the straight lengths.
b. For the semicircular arc, each infinitesimal segment, dl, is an equal distance
R from point C. For each such segment, dl×aR is directed into the page (here aR is
antiparallel to a of cylindrical coordinate system) since dl and R are perpendicular at
each point around the arc. Therefore, by the Biot-Savart Law, we have

μI dl  a R μI Rda φ (a ρ ) μI
B
4  R 2

4  R 2

4R
az
0

c. To get the net contribution, just add the results for parts a and b:
μI
B az
4R

Example 2.5: Consider the circuit shown in following figure. A current I runs through it.
Given the parameters shown in the figure, find the magnetic field at point P.

Solution:
We can take our result from the previous problem and apply it here directly. The straight
sections yield no contribution to the field at P since they lie along the radial line from
each current element to point P. We also know that a semicircular arc gives a magnetic
field of magnitude B=I/4R. An arc which is less than a semicircle gives a field of
I arc  length
B=I(arc-length)/4R or B   .
4R R
For our problem, we use the right-hand rule to note that the field due to the arc of radius
b points out of the page while the arc of radius a gives a field that points into the page.
Then, the net field at P
 I b I a  I (a  b)
B    a z  az
 4b b 4a a  4ab

Example 2.6: Find the magnetic field at point P for following figure which shows a
current I traveling a long, straight wire length into a semicircular wire of radius R and out
along another long, straight section.

Solution:
Let's consider the current distribution as consisting of 3 parts as labeled in the following
figure.
y

These pieces are the two long, straight sections and the semicircular arc. For pieces 1 and
3, note that their contributions to the net B field at point P are both out of the page (along
+ve z-axis). In setting up the Biot-Savart integral for infinitesimal lengths of current
elements dx, note that the setup of the integral will be identical for both, namely, we can
consider the position of dx in both cases w.r.t. the places where the straight sections
combine with the semi-circular arcs. Therefore, the sum of these two contributions is the
same as that of a single infinite length line carrying a current I. Hence
μI dl  a R
B1  B 3  2 
4  R2

Now dl=dx ax and R=Rcos (-ax)+ Rsin ay= -x ax+r ay, |R|=R=(x2+r2)1/2

μI dl  a R μI

dxa x  ( xa x  ra y ) / R μI

rdxa z
B1  B 3  2 
4  R 2  2 
0 R 2

2 
0 R3

μI rdxa z μI
 B1  B 3 
2  (x
0
2
r )
2 3/ 2

2 r
az
To get the contribution from piece 2, the semi-circular arc, look again at figure. In this
case, each current element is a distance r away from point P and the angle between dl, the
infinitesimal arc-length of current, and R, the vector from dl to point P, is 90 because
dl is tangential to the arc and R is along the radial direction. So,

μI dl  a R μI

rda φ  (a ρ ) μI

μI
B2 
4  r 2  4 
0 r2
 
4r 0
da z  a z
4r

Summing all the contributions together yields


μI  1 1 
B net  B1  B 2  B 3    a z
2r   2 

Example 2.7: A square conducting loop of side 2a lies on the z=0 plane and carries a
current I in the counterclockwise direction. Show that at the center of the loop
2I
H az .
a
Solution:
y
dl=dyay
x

Consider a small current element dl=dy at a height y along the right side of the square.
Then, r=0, r=aax +yay. Therefore R= r-r=–a ax - yay; R3=(a2 + y2)3/2.
The H field for the right side is then

1 I dl   R dl   R
a a
I
H right side     .
a
4 R 3
4 a R3
Now, dlR=dyay(–a ax - yay)=adyaz. Therefore,
dl   R ady 
a a
I I 2I
H right side 
4  R3

4  (a 2  y  2 ) 3 / 2 a z  4a a z
a a
The H field contributions for all four sides are same. Therefore, the resultant H field is
given by
2I
H az .
a

Example 2.8: A sheet of conducting metal carrying a dc current of IS amperes per unit
width in the z-direction as shown in the following figure. Calculate H at a point P(x0, y0,
z0).
Solution: z
R P(x0,y0, z0)
dz
z
This sheet is assumed be placed at y=0.
y
x
x dx

r=x0 ax + y0 ay + z0 az, r=xax +zaz. Therefore R= r-r=(x0 –x) ax + y0 ay + (z0 –z) az


R3={(x0 –x)2 + y02 + (z0 –z)2}3/2
 I dl   a R  I dl   R
d (dB)  
4 R2 4 R3
I=ISdx, dl=dz az
dlxR= dz az  {(x0 –x) ax + y0 ay + (z0 –z) az}= dz {(x0 –x) ay - y0 ax}

I  
{( x 0  x )a y  y 0 a x }dz dx 
B S
4  
z   x   {( x0  x ) 2  y 0  ( z 0  z ) 2 }3 / 2
2

 
I y 0 dz dx  I S IS
Bx   S
4  
z   x   {( x0  x )  y 0  ( z 0  z ) }
2 2 2 3/ 2

2
 Hx  
2
 
I ( x 0  x )dz dx 
By  S
4  
z   x   {( x0  x ) 2  y 0  ( z 0  z ) 2 }3 / 2
2
 0 H y  0

IS
H   ax
2
It can be shown similarly H=IS ax/2 for P(x0, -y0, z0).
It is very interesting to note that electrostatic field due to sheet charge was also found to
be constant like magnetostatic field, irrespective of the position of the observation point
in front of the sheet.
Example 2.9: A circular loop of radius a carries a direct current I in the anti-clockwise
direction as shown in the following figure. Calculate B at a point P on the z-axis.
Solution:
r=z az, r= a a. Therefore R=r - r=z az –a a  R3=| r-r|3= (z2+a2)3/2
dl x R= ad a  (z az –a a)= azd a + a2 d az
dl  r  r  I 2 2
μI azd a 2 d
B 
4  r  r
3

4 0 ( z 2  a 2 ) 3 / 2
[ aρ   2
0 (z  a )
2 3/ 2
az ]

The first integration over a closed circular path is zero, as a varies with . It can be
verified mathematically if we replace a by (cos  ax + sin  ay) in the first integration
and then integrate it w.r.t  from 0 to 2. The field is therefore given by
I a2 I
B a z T . At the centre of the loop B  az T
2 (z  a )
2 2 3/ 2
2a
I a 2
For z>>a B  az T .
2 z3
When the point of observation is far from the loop, the size of the loop is very small in
comparison with the distance z. In this case, we refer to the current carrying loop as a
magnetic dipole. If the magnetic dipole moment is defined as m=Ia2az, B is then given
by
m
B T
2 z 3

Example 2.10: Two circular loops of radius a carries a direct current I in the anti-
clockwise direction as shown in the following figure. Calculate |B| as a function of z/d
along the z-axis.
Solution:
Following the solution of Example 2.9, B field for coils on z=0 and z=d planes are given
by
I a2 I a2
B a and B  a z , respective ly.
2 (z 2  a 2 )3/ 2
z
 
2 (z  d )2  a 2 3/ 2

Ia 2  1 1 
Therefore the net field is B   2  3/ 2  z
 
a .
2  ( z  a )
2 3/ 2
(z  d )  a2
2

The magnitude of B field along z-axis for various values of z/d is also shown in the
following figure and is found to become maximum at z=d/2 for d=a.

Normali
zed

|B|

z/d
Such a configuration is called a Helmholtz coil.
Example 2.11: Determine the field H on the axis of a solenoid.

Solution: A solenoid is a cylindrical shaped current carrying coil. Just as the parallel
plate capacitor concentrates the electric field between the plates, the solenoid
concentrates the magnetic field with in the coil. For the purpose of determining the
solenoid magnetic field, the solenoid of length l and radius a which is tightly wound with
N turns can be modeled as an equivalent uniform surface current on the cylinder surface
as shown in the following.
The uniform surface current in A/m for the solenoid is found as
NI
K aφ .
l
Using equation (2.4a), we have
1 K  (r  r)
4  r  r 3
H ds

Now to determine the characteristics of the magnetic field inside the solenoid, we choose
the field point P on the solenoid axis as shown in the following figure.

Here, ds=addz, r=zaz, r=aa+ zaz, r-r= -aa+ (z-z)az. Therefore,

1 NIa φ  ( aa ρ  ( z  z )a z ) NIa 2


l/2
az
H
4 
S 
l a  ( z  z )
2

2 3/ 2
ad dz  
2l  a
l / 2
2
 ( z  z ) 2 
3/ 2
dz 

NI  z l /2 z l /2 
   a z
2l  a 2  ( z  l / 2) 2 2
  2

 a ( z l / 2)
The magnetic field at the center of the solenoid (z=0) is
NI  1 
H  a z
2  a 2  (l / 2) 2 
 
At either end of the solenoid (z=l/2, z=-l/2), the magnetic field is
NI  1 
H  2 2 a z
2  a l 
For a long solenoid (l>>a), the approximate magnetic field values at the center and at the
ends of the solenoid are
NI
H a z (center)
l
NI
H a z (at either end)
2l
Thus, the magnetic field at the ends of a long solenoid is approximately half that at the
center of the solenoid. However, the magnetic field over the length of a long solenoid is
relatively constant. At the ends of the long solenoid, the magnetic field falls rapidly to
about one-half of the peak value. This is shown in the following figure.

2.3 AMPERE’S CIRCUITAL LAW


Ampere‘s law is the one of the Maxwell‘s equations that relates the magnetostatic field to
its source (dc current).
Ampere‘s circuital law states that the line integral of the magnetic field intensity H
around a closed path equals the net current flowing throw the surface bounded by that
path. This statement is given mathematically by,

 H  dl  I   B  dl  μI (2.5)

The sign convention for the direction of the closed path in Ampere‘s law is taken so that I
and H satisfy the right hand thumb’s rule defined earlier in connection with Biot-Savart
law.
Equation (2.5) is the Ampere‘s law which is the special case of Biot-Savart‘s law and is
applied in problems involving symmetrical current distribution only. Ampere‘s circuital
law is very powerful in determining the magnetic field caused by a current I, when there
is a closed path around the current such that the magnitude of the field is constant over
the entire path.
By applying Stoke‘s theorem the left side of equation (2.5) can be written as

 H  dl   (  H)  ds
S
or,  B  dl   (  B)  ds
S

ds
dl

Again the total current flowing through the surface ds can be written in terms of a surface
integral of the current density J as I   J  ds
s

Hence Ampere‘s circuital law can be written in the form

  H  J or   B  J (2.6)
This result must be true for any choice of surface s or at any point in space. Equation
(2.6) is the differential form of Ampere‘s circuital law and it is the Maxwell’s third law.
It shows that the magnetic field is rotational and also not conservative.

Example 2.12: An infinitely long solid conductor of radius a is placed along the z-axis. If
I
the conductor carries a current I in the +z-direction, show that H  a within the
2a 2
conductor.
Solution:

Consider a closed circular path C (usually known as Amperian closed path) of radius 
(<a) on the conductor‘s cross-sectional surface, as shown in the following figure, and
then apply ACL along this path.

We know, the H field is circular in this case, that is, H=H a. Therefore from ACL, we
have  H  dl  I   J  ds where dl is an differential length of the path along the dotted
line and ds is an differential area on the cross-sectional area of the wire.
Therefore, from the left side of the ACL, we obtain
2

 H  dl   H  d  2H 
0
and the he right side of ACL gives
 2
I I 2
 J  ds    dd 
 0  0 a
2
a2

I 2 I I
 2H    H  H aφ .
a 2
2a 2
2a 2

Example 2.13: Six parallel aluminum wire of small, but finite radius lie in the same
plane. The wires are separated by equal distances d, and they carry equal currents I in the
same direction. Find the magnetic field at the center of the first wire. Assume that the
current in each wire is uniformly distributed over its cross section.
Solution:
A schematic layout of the problem is shown below.

Figure: Six parallel wires.


The magnetic field generated by a single wire as obtained from Ampere‘s law is equal to
μI
B aφ
2

where  is the distance from the center of the wire. The equation is correct for all points
outside the wire, and can therefore be used to determine the magnetic field generated by
wire 2, 3, 4, 5, and 6. The field at the center of wire 1, due to the current flowing in wire
1, can be determined using Ampere's law, and is equal to zero. The total magnetic field at
the center of wire 1 can be found by vector addition of the contributions of each of the six
wires. Since the direction of each of these contributions is the same, the total magnetic
field at the center of wire 1 is equal to
μI  1 1 1 1 1 
B  B1  B 2  B 3  B 4  B 5  B 6       a φ
2  d 2d 3d 4d 5d 
Example 2.14: A coaxial cable consists of a long cylindrical copper wire of radius r1
surrounded by a cylindrical shell of inner radius r2 and outer radius r3 (see Figure below).
The wire and the shell carry equal and opposite currents I uniformly distributed over their
volumes. Find the formulas for the magnetic field in each of the regions r < r1, r1 < r < r2,
r2 < r < r3, and r > r3.
Solution:
The magnetic field lines are circles, centered on the symmetry axis of the coaxial cable.
First consider an integration path for Ampere‘s Circuital Law (ACL) with r < r1. The path
(line) integral of H along this path is equal to
 r2I
 H  dl   r12 Here dl=rda and H=Ha
2
r2I r2I rI
 H  dl  r12
 
0
Hrd 
r1
2
H
2 r1
a
2 φ

For r1<r<r2, the ACL becomes


 H  dl  I Here dl=rda and H=Ha
2
I
 H  dl  I   Hrd  I  H 
0
2 r

For r2<r<r3, the ACL becomes


  r 2   r22   r32  r 2 
 H  dl  I  I  2   I  2 2  Here dl=rda and H=Ha
2 
  r3   r2   r3  r2 
2
 r32  r 2   r32  r 2  I  r32  r 2 
 H  dl  I 
 r2  r2 
 3 2 
  0 Hrd  I 
 r2  r2 
 3 2 
  H   a φ
2 r  r32  r22 
For r>r3 the ACL becomes
 H  dl  0 Here dl=rda and H=Ha
 H  dl  0  H  0a φ

Example 2.15: A closely spaced winding (toroidal winding) with N turns is wound in the
form of a ring as shown in the following figure (a). The inner and the outer radii of the
ring are a and b, respectively. The height of the ring is h. If the winding carries a current I
amps, find H within the ring and total magnetic flux enclosed by the ring.
Solution:
The current enclosed by any closed circular path of radius  where a   b is NI.
Therefore, from ACL we obtain
2

 H  d l  NI
l
  H  da
 0
φ NI

NI NI
H H aφ
2 2
NI b h
NIh
B  a φ     B  d s    B  d dza φ  lnb / a 
2 s   a z 0 2

Example 2.16: A very long hollow conductor of inner radius a and outer radius b is
located along the z axis and carries a current I in the z direction, as shown in the
following figure. If the current distribution is uniform, determine H at any point in space.

Solution:
In the region a   b, J=I/{(b2-a2)}az.
For  a, the current enclosed by any closed path is zero. Therefore, H is also zero as
obtained by ACL.
For a   b, the current enclosed by a closed circular path of radius  is

I enc   J  ds 
I
 2
 2

 a2
s

 b  a2
2
   a z  dda z 
 a  0
b 2

 a2
I

From ACL, we have


  a  I 2 2 2

 H  d l  I enc
l

b  a 
 0
 H  daφ  2 2

H 
  a  I  H    a  I a
2 2 2 2

2 b  a  2 b  a 
2 2 2 2 φ

For b, the current enclosed by a closed circular path of any radius  is I
Therefore, from ACL, we have
2
I I
 H dl  I
l


 H  da
0
φ I H
2
H
2

Example 2.17: A long cylindrical conductor of radius a carries a current of density J.
Show that the H field, inside a cylindrical cavity whose axis is displaced from that of the
conductor by a distance d, is constant and depends only on the location of the cavity and
not on its radius b.
Solution:
The cross-sectional view is shown in the adjacent figure. Assume that the current is
flowing in the z-direction. We can use superposition theorem and consider the H field
inside the cavity as that due to two long cylindrical conductors with radii b and a and
current densities J and –J, respectively. Let now find the H field at a point which is 
distance away from the centre of the conductor and  distance away from the centre of
the cavity.
y
 

a
b x
J -J

From ACL  H  dl  I   J  ds , H at the specified position due to J is given by


l s

J J Jρ
2H    2 J  H  H aφ H
2 2 2
Similarly, H at the specified position due to –J is given by
 J  J J  ρ
2 H     2 J  H    H aφ  H  
2 2 2
J  ρ J  ρ J  (ρ  ρ) J  d
Therefore total H field is H     where d is a vector
2 2 2 2
directed from the centre of the conductor to that of the cavity.

Example 2.18: H=k0(/a) a for <a, where k0 is a constant. Find J for  <a.

Solution: We know
H  J
aρ a φ az
J
1 
 




1 
z  

 2 k0 / a a z 1  2
 z
 
 k 0 / a a ρ  (2k 0 / a )a z .
0  2 k0 / a 0
2.4 GAUSS’S LAW FOR MAGNETIC FIELD
The characteristics of electrostatic and magnetostatic fields are fundamentally different
based on the existence of electric charge and the nonexistence of magnetic charge.
In electrostatic, the field lines originate from the +ve charge and terminate on the –ve
charge. However, magnetic flux lines, to which B is tangential at every point, always
close upon themselves as it is not possible to have isolated magnetic poles. This is shown
in the following figure.

Therefore, the total outward magnetic flux through any closed surface is zero i.e.,
 B  dS  0
S
(2.7)

The above equation reflects that the equivalence of an electric point charge does not exist
in nature.
Again from divergence theorem we have
 B  dS  (  B)dv .
S v

As the volume has finite value, it reveals that


  B =0 (2.8)
The equation (2.8) is referred to as the fourth Maxwell’s equation.
The divergence less property of the magnetic field indicates that the magnetic flux lines
are conservative although the magnetostatic field is not conserved as because H≠0.
Moreover, as the magnetic flux lines close upon themselves, it is not possible to have
isolated magnetic poles. Therefore, magnetostatic field is an example of solenoidal field.
Example 2.19: For the magnetic field of to a long current carrying conductor, show that
  B =0.

Solution:
The B field of a long conductor is given by Biot-Savart law as:
μI dl  a R
B
4  R2
μI dl  a R
Then   B 
4  
R2
.

Now R can be assumed as R=|r-r|=(x-x)ax+(y-y)ay+(z-z)az, where (x,y,z) and (x, y,


z) are respectively the coordinates of the observation and source points, and r and r are,
respectively, the position vectors of the observation and source points. Then,
1
( )  {(x - x) 2  (y - y) 2  (z - z) 2 }1 / 2
R
a
 x - x{ }3 / 2 a x  y - y{ }3 / 2 a y  z - z{ }3 / 2 a z   3   R2
R
R R
μI dl  a R μI 1
  B 
4  
R 2

4    dl  ( )
R
With the help of the vector identity AB=BA - AB the above equation can be
rearrange as
μI 1 1
B  
4  [( )    dl  dl    ( )] .
R R
1
However from vector identity,   ( )=0. The first term of the above equation is also
R
equals to zero because  is function of (x, y, z) and dl is function of (x, y, z).
Therefore,   B =0.

2.5 MAGNETIC SCALAR AND VECTOR POTENTIALS


Like electrostatic field, we can define a potential associated with magnetostatic field. In
fact, the magnetic potential could be scalar Vm or vector A. The definition of Vm and A
involves the following two vector identities:
  (V )  0 (2.9)
  (  A)  0 (2.10)
which must always hold for any scalar field V and vector field A.
Just as E=-V, we define the magnetic scalar potential Vm (in amperes) as related to H
according to
H= - Vm if J=0 (2.11)
The condition attached to this equation, J=0, comes from the Ampere‘s circuital law
which suggests that the magnetic scalar potential Vm is only defined in a region where
J=0,i.e., in a source free region.
Again since   B =0 and from the vector identity given in equation (2.10), we can define
B=   A where A be a vector, called the vector magnetic potential. The unit of the
vector magnetic potential A is Weber/m.
Derivation of vector magnetic potential due to current element

With the help of the previous section, we have

μI dl  a R μ Jdv  a R μ 1 μ 1
B
4  R 2

4  R 2

4  Jdv  ( ) 
R 4  ( )  Jdv
R

From the vector identity fF=fF + fF we have

μ 1 1
B
4  [  ( Jdv) - ( )  Jdv]
R R
The second term is zero, because  is a function of (x,y,z) and J is a function of (x, y,
z).
μ 1 μ 1
B 
4    ( Jdv)    [
R 
4 R
( Jdv)]
μ 1 μ 1 μ 1
 A(r ) 
4 R J (r )dv( r )  
4 R
I(r )dl (r )  
4 r - r 
I(r )dl (r ) (2.12)

Divergence of vector magnetic potential

Let us now consider the divergence of A. We have from eq. (2.12)


μ 1
  A(r ) 
4    J (r )dv( r )
R
Using the vector identity   ( fF)  f  F  F  f the above equation takes the form
μ 1 1
  A(r ) 
4  [J (r )  ( )    J (r )]dv( r )
R R
(2.13)

The second term of eq.(52) is zero; therefore eq.(52) becomes


μ 1 μ 1
  A(r ) 
4  [J (r )  ( )]dv( r )    A(r )  
R 4  [J (r )  ( )]dv( r )
R
1 1
because     .
R R
μ 1 J (r )
  A(r )  
4 R
[   J (r )   
R
)]dv( r ) From continuity equation the 1st term is

zero.
μ J (r ) μ J (r )
   A(r )  
4   
R
dv( r )  
4  R
 ds(r )

As the surface s(r) must include all currents, there will be no current through the surface
s(r). Therefore
J  ds(r )  0    A(r )  0 (2.14)

Vector magnetic potential in terms of current density


From ACL, we have
  H  J    B  J
As B=A, ACL yields
    A  J  (  A)   2 A  J
But, ·A=0
  2 A  J (2.15)

Special note: Except for simple current distributions with symmetrical geometries that
lend themselves to the application of ampere’s circuital law, in practice we often use the
approaches provided by the Biot-Savart law and the vector magnetic potential, and
among these two the latter often is more convenient to apply because it is easier to
perform.

2.6 MAGNETIC DIPOLE


When the magnetic field of a small circulating current loop is measured at a far point
from the loop, then we refer to the current carrying loop as a magnetic dipole. Here, in
this section, we derive the equations of vector magnetic potential and magnetic flux
density from a magnetic dipole.

Let a be the radius of a current circulating loop where r>>a. The coordinate of the
observation point P and source Idl′=Iad′a′ in rectangular coordinate system are,
respectively. (x=rsincos, y=rsinsin, z=rcos) and (x′=acos′, y′=asin′, z′=0). The A
field at P is then given by
 0 I dl   0 I ad 
4  R  ( x  x)a
A  a φ
4  
x  ( y  y )a y  ( z  z )a z

Now ( x  x)a x  ( y  y )a y  ( z  z )a z


1

 ( x  x) 2  ( y  y ) 2  ( z  z ) 2 
1 / 2

 a sin 

 r 2  a 2  2ra sin  (cos  cos    sin  sin  ) 
1 / 2
 r 1 1 

(cos  cos    sin  sin  )
 r 
again a φ   sin  a x  cos  a y
0 I ad 
A 
4  ( x  x)a  
x  ( y  y )a y  ( z  z )a z
a φ

 0 Ia  a sin  

4r   1
r
(cos  cos    sin  sin  ) ( sin  a x  cos  a y )d 

 0 Ia 2  a sin  
4r 0 
 1 (cos  cos    sin  sin  ) ( sin  a x  cos  a y )d 
r 
 0 Ia 2 sin  2

4r 2 0 (cos cos   sin  sin  )( sin  a x  cos a y )d 
 0 Ia 2 sin 
 ( sin a x  cos a y )
4r 2
 0 Ia 2 sin   0 Ia 2 sin 
Ar  2
( sin a x  cos a y )  a r  ( sin a x  cos a y )  (sin a ρ  cosa z )  0
4r 4r 2
 Ia 2 sin   0 Ia 2 sin 
A  0 (  sin a x  cos a y )  a θ  ( sin a x  cos a y )  (cosa ρ  sin a z )  0
4r 2 4r 2
 Ia 2 sin   0 Ia 2 sin 
A  0 (  sin a x  cos a y )  a φ  ( sin a x  cos a y )  ( sin a x  cos a y )
4r 2 4r 2
 Ia 2 sin 
 0
4r 2
 0 Ia 2 sin 
A  a φ Wb/m (2.16)
4r 2

ar ra θ  2  0 Ia 2 sin  cos  
r sin a φ
1 1  ar 
Now B    A  2     2  4r 
r sin  r   r sin    0 Ia 2 sin 2  
 0 Ia 2 sin 2   4 r
aθ 
0 0
4r
0 Ia 2
B  2 cosar  sin aθ  T (2.17)
4r 3
Equation (2.16) can be rearranged as
 0 Ia 2 sin   0 Ia 2 sin   ma
A aφ  a φ  0 2 r Wb/m (2.18)
4r 2 4r 2
4r
where m=Ia2az=maz Am2 is defined as the magnetic dipole moment. It is a vector
quantity whose magnitude is the product of the current in and the area of the loop and
whose direction is the direction of the thumb as the other fingers of the right hand follow
the direction of the current.
In a similar manner equation (2.17) can be rewritten as
0 m
B 2 cos a r  sin aθ  T (2.19)
4r 3
The magnetic field pattern of a magnetic dipole is shown in the following.

2.7 CALCULATION OF MAGNETIC FLUX


The magnetic flux  (measured in weber) for a given current distribution can be obtained
either from known magnetic flux density B or from known vector magnetic potential A
using the following equation:

   B  ds   (  A)  ds   A  dl (2.20)

Example 2.20: A direct current I flows in a straight wire of length 2L. Find the magnetic
flux density B at a point located at a distance  from the wire on the bisecting plane by
determining the vector magnetic potential A first.

Solution: At a height z along the wire above the bisecting plane assume a current
element Idl= Idz az. A is then given by
 Idz  I  L2   2  L 
L
A
4  z  
az 

ln
 
az .
 L   L
z  L
2 2 4 2 2

B is obtained from
aρ a φ az
1       I  L2   2  L 
B  A   Az a φ    ln a
   z    4  L    L 
 φ
2 2
 
A A Az
IL
 aφ
2  L2   2

Example 2.21: A very long straight conductor located along z axis carries a current I in
the z direction. Obtain an expression of A at a point on the bisecting plane of the
conductor. Also determine B at that point.
Solution: The conductor is assumed to be extending from z=-L to z=L as shown in the
following.

The vector R from the current element to the field point is given by R= a - z az. Now
A at P is given by
μ 1
A
4  R Idl  where dl   dz  a z and R   2  z  2

dz  a z
L
μI
A 
4 
 2  z2
L

Assume z=tan  dz=sec2 d and (2 +z2)1/2=sec


dz   2  z2  z
   sec  d  ln(sec   tan  )  ln( )
 2  z2 

I  2  L2  L
Therefore A  ln az
4  2  L2  L
2 2 2
But  2  L2  L  L  L[1  ]1/ 2  L  L[1  2 ]  2 L   2L
L2 2L 2L
 2 1/ 2 2 2
and   L  L   L  L[1  2 ]   L  L[1  2 ] 
2 2

L 2L 2L
I 2 L I 2 L
A  ln 2 a z  ln az
4  2 
2L
Now as B=A

aρ a φ az aρ a φ az
1    1     I
B    A    Az a φ  aφ
   z    z  2
A A Az 0 0 Az

Example 2.22: The inner conductor of a 100m long coaxial cable has a radius of 1cm
and carries a current of 80 A in the z direction as shown in the following figure. The
outer cylinder is very thin and has radius of 10 cm. Calculate the total flux enclosed
within the conductors.

Solution: We can use (2.20) and A or (2.20) and B to determine flux in the following
way:
   A  dl or    B  ds
If we use A, then

As A has only z-component, then we have

Again from
I b L
ddz IL b I
   B  ds     ln because here ds  ddza φ and B  aφ
2 
a z  L
  a 2

Example 2.23: If B=Baz, compute the magnetic flux passing through a hemisphere of
radius R centered at the origin and bounded by the plane z-0.

Solution:
We know    B  ds where ds=r2 sindd. In this problem B=Baz and r=R.
 / 2 2
   B  ds    Ba z  R 2 sin dda r
 
0 0

Now azar=cos

 / 2 2  /2  /2
    Ba  R sin dda r  2R B  sin  cos d  R B  sin 2d  R 2 B
2 2 2
z
 
0 0  0  0

Example 2.24: Two very long identical and parallel conductors carrying 1 kA in
opposite directions are strung on poles 100m apart. If the radius of each conductor is 2cm
and the separation between their axes is 1m, determine the flux passing through the
region bounded by the conductors and the two consecutive poles.
Solution:
Here a=2cm, b=1m, L=100m, I=1kA. Magnetic flux density B at a distance y from the
conductor at y=0 is

I 1 1 
B  y  b  y  (a x )
2  
and    B  ds where ds  -dydza x
L ba
I 1 1  IL  b  a 
   
z 0 y  a
2  y  b  y  dydz   ln  a   155.67 mWb
   

Example 2.25: Determine the magnetic flux through a rectangular loop (ab) due to an
infinitely long conductor carrying current I as shown in the following figure.
Solution: The magnetic flux density at a perpendicular distance  from an infinitely long
current carrying conductor is given by
I
B a .
2
Now from equation (2.20)
d  a z b
I Ib d  a
   B  ds    a φ  ddza φ  ln
s   d z 0
2 2 d

Example 2.26: A brass ring with triangular cross section encircles a very long straight
wire concentrically as shown in the following figure. If the wire carries a current I, show
that the total number of magnetic flux lines in the ring is
Ih ab
  (b  a ln )
2b a

Solution: On the slant side of the ring, z=(-a)h/b.


a b z  h (   a ) / b a b
I Ih (   a) Ih a b
    B  ds    a φ  ddza φ   d  (   a ln  ) a
s  a z 0
2 2b  a  2b
Ih ab
 (b  a ln )
2b a

Example 2.27: Find the current density J due to


10
A  2 a z in free space.

Solution: From the Poisson‘s equation of A in (2.15) we have
 2 Az    0 J z
1   Az  1  2 Az  2 Az
   2    0 J z
      
2
z 2
For the given A,

1   Az 
    0 J z
    
1     10   1   20  40
   2      0 J z    2     0 J z  4    0 J z

             
40
 Jz  
0 
4

40
J   az ( A / m2 )
0  4

Example 2.28: A coaxial transmission line is constructed such that the radius of the inner
conductor is a and the outer conductor has radii 3a and 4a. Find the vector magnetic
potential within the outer conductor. Assume Az=0 for =3a.

Solution: For the outer conductor


I I
Jz   
 (16a  9a )
2 2
7a 2
Now from Poisson‘s equation with above current density
1   Az  1  2 Az  2 Az
   2    0 J z
      
2
z 2
1   Az  I
    0
     7a 2
  Az  I
     0
    7a 2
Integrating the above equation once w.r.t , we obtain
Az I 2
  0  c1
 14a 2
Again, integrating the above equation w.r.t , we obtain
I 2
Az   0  c1 ln   c 2
28a 2
Using the boundary condition that Az=0 at =3a, we get
9 Ia 2
0  0  c1 ln 3a  c 2
28a 2
9I
 c 2  c1 ln 3a   0
28
I  2  
 Az   0   9   c1 ln

28  a 2 
 3a
To determine the constant c1, let us use

  A  B  0 H
a a az
1      I  2  
 LHS   0  2  9   c1 ln a φ
   z   28  a  3a 
A A Az
 I  c1 
  0  a φ
 14 a 
2

Now from ACL, at =3a


I
 H  dl  I  2 3aH   I  H   6a
 I 3a c1  I  I I 3 8 0 I
  0      c  3a 
 6a    
 14 a 6a 14 a  7
0 1 0 0
2
3a  
Therefore,
I  2  
Az   0   9   c1 ln
28 a 2 
  3a
I  2  8 0 I 
 0   9  ln
28  a2  7
  3a

2.8 LORENTZ FORCE


In Chapter 1, we have discussed about the electric force on a stationary charge Q in an
electric field E given by
Fe  QE. (2.21)
A moving charge in the electric field also experiences the same force as given by
equation (2.21). However, a magnetic field can exert force only on a moving charged
particle. From experiments, it is found that the magnetic force Fm experienced by a
charge Q moving with a velocity u in a magnetic field of flux density B is
Fm  Q(u  B). (2.22)
The electric force Fe is found to act parallel to the direction of electric field; the magnetic
force Fm, however, is perpendicular to both u and B.
From equations (2.21) and (2.22), a comparison between the electric force Fe and the
magnetic force Fm can be made. Fe is independent of the velocity of the charge and can
perform work on the charge and change its kinetic energy. Unlike Fe, Fm depends on the
charge velocity and is normal to it. Moreover, the magnetic force Fm cannot perform any
work because it is at right angles to the direction of motion of the charge (dW mag= Fm·dl=
Fm·udt=Q(uB)·udt= 0). The magnetic force can not change the kinetic energy of a
charge particle; the magnetic force can change the direction of motion of the charge
particle, but not its speed. On the other hand, the magnitude of Fm is generally small in
comparison to Fe except at high velocities.
For a moving charge in the presence of both electric and magnetic fields, the total force
on the charge is given by
F  Fe  Fm
or, F  Q(E  u  B). (2.23)
This is known as Lorentz force equation. It relates mechanical force to electromagnetic
force. If the mass of the charged particle moving in E and B fields is m, we have from
Newton‘s second law of motion
du
Fm  Q(E  u  B). (2.24)
dt
The solution to this equation is important in determining the motion of the charged
particles in E and B fields.

2.8.1 Magnetic Force on a Current Carrying Conductor


Assume an element of current, as a length dl of wire carrying a current I. The cross-
sectional area of the wire and the density of charge in the wire are considered S and v
respectively. Assume the charges move with velocity u. Then the charge crossing any
cross-section in one second=uSv=I.
Idl= uSdlv=uq.
Therefore, the force on the differential current element dl due to magnetic field only can
be expressed as
dF=quxB=IdlxB. (2.25)
A current carrying conductor placed in a magnetic field experience a force that tends to
move the conductor in a direction perpendicular to both magnetic field and the conductor.
For a real circuit L, the total magnetic force is given by the appropriate integration of the
forces acting on the individual current elements:
F   Idl  B (2.26)
L

Equation (2.26) is known as Ampere’s force law.


If a closed current carrying wire resides in a uniform external magnetic field B as shown
in the following figure, then B can be taken outside the integral in equation (2.26), in
which case
 
F  I   Idl   B  0 .
L 
This result, which is a consequence of the fact the vector sum infinitesimal vectors dl over
a closed path equals zero, states that the total magnetic force on any closed current loop
in a uniform magnetic field is zero.

2.8.2 Magnetic Torque on a Current Carrying Loop


When a force is applied on a rigid body that can pivot about a fixed axis, the body react
by rotating about that axis. The angular acceleration depends on the cross product of the
applied force vector F and the distance vector r, measured from a point on the rotation
axis, such that r is perpendicular to the axis, to the point of application of F. In the
following figure the force F acting on a circular disk that can pivot along the z-axis
generates a torque =rF that causes the disk to rotate. The unit of torque is Nm.

The force F applied on the disk in the above figure lies in the x-y plane and makes an
angle  with r. Hence,
=rF=rFsin az. (2.27)
From equation (2.27), we observe that a torque along the positive z-direction corresponds
to a tendency for the cylinder to rotate counterclockwise and, conversely, a torque along
–z-direction corresponds to clockwise rotation.
We will now consider, in the following sub-articles, the magnetic torque exerted on a
conducting coil under the influence of magnetic force. Firstly, we begin with the simple
case where the magnetic field B is parallel to the plane of the coil as shown in the
following figure. Here the coil is taken in the z-x plane.

Under the action of an externally applied uniform magnetic field B=B0ax, arms ab and cd
of the coil are subjected to forces Fab and Fcd, are given by
Fab=I(baz)(B0ax)= IbB0ay
and
Fcd=I(baz)(B0ax)=IbB0 ay
No magnetic force is exerted on either arm bc or da because B is parallel to the direction
of the current element in those arms.
A bottom view of the coil, depicted in Fig. (b), reveals that forces Fab and Fcd produce a
torque about the origin O, causing the coil to rotate in the clockwise direction. The net
torque on the coil is

τ   aˆ x    IbB0 aˆ y     aˆ x   IbB0 aˆ y    IabB0 aˆ z  m  B Nm


a   a 
2   2 
where m=Iabay is the magnetic dipole moment.
As soon as the coil starts to rotate, the torque decreases, and at the end of one quarter of a
complete rotation, the torque becomes zero as discussed in the following sub-article.

Magnetic field perpendicular to the plane of the loop

We will extend the above analysis to the more general case where the plane of the coil
makes an angle  with the y-axis as shown in the following figure. In this case, all the
four arms have nonzero forces.

(a) (b)

The forces experienced by the sides ab and cd are still the same as before. However, the
magnetic force on sides bc and da are given by
Fbc=I(asinax +a cosay)(B0ax)= IacosB0az
and
Fda=I(asinax a cosay)(B0ax)= IacosB0az
Since the line of action of the forces Fbc and Fda is the same, the resultant force along z-
axis is zero. Thus, the only forces that contribute to the torque are Fab and Fcd as shown
in the above Fig. (b). Now, the torques acting on sides ab and cd are
τ ab 
a
sin aˆ x  cosaˆ y   IbB0 aˆ y    1 IabB0 sin aˆ z
2 2
and
τ cd 
a
 sin aˆ x  cos aˆ y  IbB0 aˆ y    1 IabB0 sin aˆ z
2 2
The resultant torque is
τ  τ ab  τ cd   IabB0 sin aˆ z  m  B (2.28)
As the torque experienced by the coils varies sinusoidally, the torque becomes maximum
when the plane of the coil is parallel to the magnetic field and it becomes zero when the
plane of the coil is normal to the magnetic field. If the plane of the coil is not normal to
the magnetic field, the direction of the torque is such that it tends to rotate the coil until
its plane becomes normal to the field. In other words, m tends to align with B.

2.9 MAGNETIC FORCE BETWEEN CURRENT


CARRYING CONDUCTORS
When two current carrying conductors are brought into close proximity, each conductor
lies in the magnetic field of the other conductor. Therefore, both current carrying
conductors exert a force on the opposite conductor.
Let us consider two current elements I1dl1 and I2dl2 are separated by a distance R as
shown in the following figure:
z
R
aR R I2dl2
I1dl1 r r
y

The magnetic field at the position of the current element I1dl1 due to the current element
I2dl2 is found from equation (2.2) as
 I 2 dl 2  a R R r  r
dB  where a R   where r and r are the position vectors of
4 R 2
R r  r
the current elements I1dl1 and I2dl2, respectively.
Following equation (2.25), the force d(dF1) on the differential current element I1dl1, due
to the differential current element I2dl2 is therefore given by
d(dF1)= I1 dl1 x dB
μI dl  (I 2 dl 2  a R )
 d(dF1 )  1 1
4R 2
Total force on circuit L1, due to the circuit L2 is then given by
μI1 I 2 (dl  a )
F1 
4 L1  [dl 1   2 2 R ]
L2 R
Newton (2.29)

Using the vector identity A(BC)= B(A·C)- C(A·B), we obtain from (2.29)
μI1I 2 dl1  (dl 2  R ) μI1I 2  dl1  R dl1  dl 2 
F1 
4  [ R3
]  
4  L L R 3
dl 2    R 3 R 
L1 L2 1 2 L1 L2 
As because R/R3= -(1/R), the first integral on the right side of the above equation can
be writen as follows with the help of Stock‘s theorem:
 1   1 
   ( )  dl1  dl 2       ( )  ds1 dl 2
L1 L2   L2 s1  
R R

The above integral is equal to zero as the curl of a gradient of a scalar function is zero.
Thus the magnetic force on circuit 1 due to current in circuit 2 is also given by
μI1I 2 dl1  dl 2 μI I dl1  dl 2 μI I dl1  dl 2
F1  
4 
L1 L2
R 3
R 1 2
4 
L1 L2
R 2
aR   1 2
4 
L1 L2 r  r 
3
(r  r) (2.30)

Therefore, the force is attractive if the currents flow in the same direction, and the force is
repulsive if the currents flow in opposite directions.
Example 2.29: A charged particle of mass m kg and charge of Q C with an initial
velocity u=u0 ay m/s enters into a region of magnetic flux density B= B0 az Wb/m2. Find
the velocity of the particle into the region.
Solution: The force balancing equation is
d
m u  Q(u  B)
dt
d
dt
   
m u x a x  u y a y  u z a z  Q u x a x  u y a y  u z a z  B0a z 
d QB0
ux  u y  c u y
dt m
d
u y  c u x .
dt
d
uz  0
dt
In the above equations, c is the cyclotron frequency.
Now, the solutions for the velocity functions are
u y  A cos  c t  B sin  c t
u x  A sin  c t  B cos  c t
uz  C
Using the initial value we have
u y  u0 cos c t  y  (u0 /c ) sin c t
u x  u0 sin c t  x  (u0 /c ) cos c t
uz  0  z0
2
u m
 x  y  (u0 /c )   0 
2 2 2

 QB0 
The above results show that the charged particle gives a circular motion in the constant
magnetic field; and the radius of the circle increases with the increase of tangential
velocity u.

Example 2.30: An electron with velocity u=(3ax+12ay-4az)105m/s experiences no net


force at a point in a magnetic field B=10ax+20ay+30az mWb/m2. Find E at that point.

Solution: The force on an electron is


F  e(E  u  B) .
Now according to the problem statement

 e(E  u  B)  0  E  (u  B)
ax ay az
E 3 12  4  (44a x  13a y  6a z )kV / m
10 20 30
Example 2.31: A particle with a mass of 1 kg and a charge of 2C starts from rest at (2, 3,
-4) in a region where E=-4 ay V/m and B=5 ax Wb/m2. Calculate
(a) The location of the particle at t=1 s
(b) Its velocity and KE at that location.
Solution: From the force balancing equation for this problem, we obtain
du
m  Q (E  u  B )
dt
d
 (u x a x  u y a y  u z a z )  2[4a y  (u x a x  u y a y  u z a z )  5a x ]
dt
d
ux  0
dt
d
u y  8  10u z
dt
d
u z  10u y
dt
d2
 2 u y  100u y  0
dt
 u y  A cos 10t  B sin 10t
u z  0.8  A sin 10t  B cos 10t
ux  C
At t  0 u  0
 C  0, A  0, and B  -0.8.
 u  (0,  0.8 sin 10t , 0.8  0.8 cos 10t )
Again, x  c1
y  0.08 cos10t  c2
z  0.8t  0.08 sin 10t  c3
At t  0, (x, y, z)  (2, 3, - 4)
 c1  2, c 2  2.92, c3  4
x  2
y  0.08 cos 10t  2.92
z  0.8t  0.08 sin 10t  4
(a ) at t  1s, (x, y, z)  (2, 2.853, - 3.156)
(b) Displacement S  2a x  ( 0.08 cos 10t  2.92 )a y  ( 0.8  0.08 sin 10t  4 )a z
d
 u(at t  1 s)  [2a x  ( 0.08 cos 10t  2.92 )a y  ( 0.8  0.08 sin 10t  4 )a z ]
dt
 0.435a y  1.471a z m/s
1
KE  m u  1.177 J
2

Example 2.32: A bent wire as shown in the following figure lies in the xy plane and
carries a current I. If the magnetic flux density in the region B=B az, determine the
magnetic force acting on the wire.

Solution: Let F1, F2 and F3 are the magnetic forces act on the sections of the wire from
x=-(a+L) to x=-a, from x=a to x=(a+L), and on the semicircular section of radius a,
respectively. Now
a a
F1   Idxa x  Ba z  
( a  L )
 BIdxa
( a  L )
y   BILa y
(a L) (a L)

F2   Idxa x  Ba z  
a
 BIdxa
a
y   BILa y

0 0 0 0
and F3   Idl  Ba z   Iada φ  Ba z   IBad (a ρ )   IBa (cos a x  sin a y )d
   

 2IBaa y
 F  F1  F2  F3  2 BI (a  L)a y

Example 2.33: The following figure shows a current-carrying conductor of finite length
L placed at a distance b from another current carrying conductor of infinite extent.
Determine the magnetic force per unit length acting on the finite conductor.

Solution: We have B at a distance b from a current carrying conductor of infinite extent


is
 I
B  0 a φ . Now from (2.25), the force on element dz is given by
2b
0 I 0 I 2
dF  Idl  B  Idz a z  aφ  dza ρ
2b 2b
L/2
 I2  I2
 F   0 dza ρ  0 La ρ
L / 2
2b 2b
0 I 2
Now the force per unit length  aρ N / m
2b
Example 2.34: A rectangular loop carrying current I2 is placed close to a straight
conductor carrying current I1, as shown in the following figure. Obtain an expression for
the magnetic force experienced by the loop.
Solution: The total magnetic force acting on the loop can be expressed as the sum of the
forces on sections AB, BC, CD, and DA. The incremental segments for sections AB and
CD are along z-axis, therefore, their dot product with the segment dz1 exist. However, the
segment dz2 for sections BC and DA is perpendicular to dz1, thus their dot product with
the segment dz1 is zero do not contribute to the force on the loop.
Let us first determine the magnetic force on section AB. The vector R is given by R=bay
- (z1 – z2)az. From (2.24), we obtain the force on section AB of the loop is

μI1I 2
L a
[ba y  ( z1  z 2 )a z ]
FAB  
4 
z1   L
dz1 
z2   a
[b 2  ( z1  z 2 ) 2 ]3 / 2
dz2

μI1I 2
FAB   I int 1  I int 2 
4
L a
bdz2
where I int1  
z1   L
dz1 
z2   a
[b  ( z1  z 2 ) 2 ]3 / 2
2
ay

( z1  z 2 )dz2
L a
and I int2  
z1   L
dz1 
z2   a
[b  ( z1  z 2 ) 2 ]3 / 2
2
az

To solve Iint2, let us consider z1- z2= b tan , therefore dz2= -b sec2 d.
a
sin d
L a L
dz1
 I int2    dz1  az   az
z1   L z2   a
b z1   L
[b  ( z1  z 2 ) 2 ]1/ 2
2
z2   a

a L

 ln[(( z1  z 2 )  b 2  ( z1  z 2 ) 2 ] az  0
z2   a z   L
1

To solve Iint1, let us consider now z1- z2= b tan , therefore dz2= -b sec2 d.
a
cos d ( z1  z 2 )dz1
L a L
 I int1    dz1  ay    ay
z1   L z2   a
b z1   L
b[b  ( z1  z 2 ) 2 ]1/ 2
2
z2   a

Put again z1- z2= b tan , therefore dz1=b sec  d. 2

a L
a
L
b  ( z1  z 2 )
2 2
 I int1    tan  sec da
z1   L
y 
b
ay
z2   a z2   a z   L
1

b 2  ( L  a) 2  b 2  ( L  a) 2
2 ay
b
μI1I 2  b  ( L  a)  b  ( L  a) 
2 2 2 2
μI1I 2
 FAB   I int 1  I int 2     ay 
4 2  b 

The force FAB is attractive in nature as the term within the bracket is positive.
Let us now determine the magnetic force on section CD. The vector R is given by R=cay
- (z1 – z2)az. From (2.29), we obtain the force on section CD of the loop is
μI I
L a
[ca y  ( z1  z 2 )a z ]
FCD  12
4 
z1   L
dz1 
z2   a
[c 2  ( z1  z 2 ) 2 ]3 / 2
dz2

μI1I 2
FCD   I int 1  I int 2 
4
L a
cdz2
where I int1  
z1   L
dz1 
z2   a
[c  ( z1  z 2 ) 2 ]3 / 2
2
ay

( z1  z 2 )dz2
L a
and I int2  
z1   L
dz1 
z2   a
[c  ( z1  z 2 ) 2 ]3 / 2
2
az

To solve Iint2, let us consider z1- z2= c tan , therefore dz2= -c sec2 d.
a
sin d
L a L
dz1
 I int2    dz1  az   az
z1   L z2   a
c z1   L
[c  ( z1  z 2 ) 2 ]1/ 2
2
z2   a

a L

 ln[(( z1  z 2 )  c 2  ( z1  z 2 ) 2 ] az  0
z2   a z   L
1

To solve Iint1, let us consider now z1- z2= c tan , therefore dz2= -c sec2 d.
a
cos d ( z1  z 2 )dz1
L a L
 I int1    dz1  ay    ay
z1   L z2   a
c z1  L
c[c  ( z1  z 2 ) 2 ]1/ 2
2
z2   a

Put again z1- z2= c tan , therefore dz1=c sec  d. 2

a L
a
L
c  ( z1  z 2 )
2 2
 I int1    tan  secda
z1   L
y 
c
ay
z2   a z2   a z   L
1

c 2  ( L  a) 2  c 2  ( L  a) 2
2 ay
c
μI1I 2 μI1I 2  c 2  ( L  a) 2  c 2  ( L  a) 2 
 FCD  I int 1  I int 2    ay 
4 2  c 

The force FCD is repulsive as the term in the bracket is positive. The total force on the
loop is therefore given by
F  FAB  FCD

μI 1 I 2  b  ( L  a )  b  ( L  a ) c 2  ( L  a) 2  c 2  ( L  a) 2 
2 2 2 2
   a y
2  b c 

As because c>b, the force on the loop is attractive.
Example 2.35: A circular coil of 200 turns has a mean area of 10 cm2, and the plane of
the coil makes an angle of 300 with the uniform magnetic flux density of 1.2 T, as shown
in the following figure. Determine the torque experienced by the coil if it carries a current
of 59 A.
Solution:

Coil in B field Side view of the coil

The side view of the coil indicating the direction of the dipole moment is shown in Figure
(b). It lies in the xy-plane and has magnitude of m=NIA=200501010-4=10 Atm2. The
torque experienced by the coil is
τ  m  B  10(sin 60 a x  cos 60 a y )  1.2a y  10.39a z N  m.

2.10 MAGNETIZATION IN MATERIALS


Magnetization in a material is due to atomic scale current loops associated with: (1)
orbital motions of the electrons and protons around and inside the nucleus and (2)
electron spin. The magnetic moment due to proton motion typically is three orders of
magnitude smaller than that of the electrons, and therefore the total orbital and spin
magnetic moment of an atom is determined by the sum of the magnetic moments of its
electrons. Both these electronic motions produce internal magnetic fields Bi that are
similar to the magnetic field produced by a current loop discussed in Art. 2.6. The
equivalent current loop, as produced due to orbital motion and spinning of atomic
electrons, has a magnetic moment of m=IbSan, where S is the area of the loop and Ib is the
bounded to the atom.
In absence of externally applied B field to a material, the sum of m‘s is zero due to
random orientations as shown in Fig. (a) in the following.

When an external magnetic field B is applied, the magnetic moments of the electrons
more or less align themselves with B so that the net magnetic moment is not zero, as
illustrated in Fig. (b).
Now, the magnetization M, in amperes per meter, is defined as the number of magnetic
dipole moment per unit volume in a material which is proportional to the H in case of
linear and isotropic material and is given by M=mH, where m is a dimensionless
constant known as the magnetic susceptibility. The net magnetic flux density, therefore,
given by B=0(H+M)= 0(H+mH)= 0(1+m)H=0rH=H, where  and r are,
respectively, the permeability and relative permeability of the material. The unit of
permeability is henrys/m.

2.10.1 Classification of Materials

In general, magnetic susceptibility m or the relative permeability r of materials are used


to classify them in terms of their magnetic property. A material is said to be nonmagnetic
is m=0 or r=1; it is magnetic otherwise. Free space, air, and materials with m=0 or
r=1 are regarded as nonmagnetic.
Materials usually are grouped into three major classes: diamagnetic, paramagnetic, and
ferromagnetic. Diamagnetic materials have slightly negative susceptibilities (r≤1)
whereas paramagnetic materials have very small positive susceptibilities (r≥1). Thus,
r≈1 or ≈0 for diamagnetic and paramagnetic substances, which include dielectric
materials and most metals. In contrast, r >> 1 for ferromagnetic materials; r of purified
iron , for example, is on the order of 2105.
Diamagnetism occurs when the magnetic fields in a material due to electronic orbital
motions and spinning completely cancel each other; the magnetic moments of each atom
is, therefore, zero and such materials are weakly affected by an external magnetic field.
For most diamagnetic materials such as bismuth, lead, copper, silicon, diamond, sodium
chloride, gold, mercury, silver etc., m is of the order of -10-5. In superconductors, the
perfect diamagnetism occurs at temperature near absolute zero: m=-1 or r=0 and B=0.
Thus superconductors cannot contain magnetic fields.
Paramagnetism occurs when the magnetic fields in a material due to electronic orbital
motions and spinning do not cancel completely. Unlike diamagnetism, Paramagnetism is
temperature dependent. For most diamagnetic materials such as air, platinum, tungsten,
potassium, aluminum, calcium, chromium, magnesium, niobium etc., m is of the order +
10-5 to +10-3.
Ferromagnetism is the property of iron, cobalt, nickel, and the compounds and alloys of
these elements. Some of the electrons in these materials have their resultant magnetic
dipole moments aligned, which produces regions with strong magnetic dipole moments.
An external magnetic field can then align the magnetic moments of such regions,
producing a strong magnetic field for a sample of the material; the field partially persists
even when the externally applied magnetic field is removed. Ferromagnetic materials are
very useful in practice. Hard ferromagnetic materials are characterized by wide hysteresis
loops and cannot be easily demagnetized. They are used in the fabrication of
electromagnets for electric motors and generators where residual magnetism is of prime
importance. However, the soft magnetic materials have narrow hysteresis loops and
hence can be more easily magnetized and demagnetized. They are used to fabricate
transforms where residual magnetism is not required.

2.11 BOUNDARY CONDITIONS FOR MAGNETOSTATIC


FIELD

The fundamental boundary conditions involving magnetic fields relate the tangential
components of magnetic field and the normal components of magnetic flux density on
either side of the media interface. The same techniques used to determine the electric
field boundary conditions can be used to determine the magnetic field boundary
conditions. The tangential magnetic field boundary condition is found by applying
Ampere‘s law on a closed path that straddles (overlaps) the media interface while the
normal magnetic flux boundary condition is found by applying Gauss‘s law for magnetic
fields to a volume straddling the media interface. The resulting boundary conditions are
shown below:
Tangential magnetic field

The boundary condition for the tangential components of magnetic field is obtained from
 H  dl  I
L
Let us take the rectangular path as the closed loop. If its height y is made infinitesimally
small then the only contribution to the line integral is along the top and bottom edges of
length x. A sheet of current having a linear current density of K A/m width is also
assumed at the interface of the two media.

Then  H  dl  H
L
t2  (xa x )  H t1  (xa x )  H t1x  H t2 x  Kx

leads to Ht2-Ht1=K (2.31)


The tangential component of H is continuous (K=0) across the boundary of almost all
physical media. It is discontinuous (K0) only when an interface with an ideal perfect
conductor is assumed because there will be no magnetic field inside the perfect
conductor.

Normal magnetic flux density

For the normal field components, let us assume a small closed box with its top face in
medium 1 and bottom face in medium 2. The faces have area dS=xy, and the height of
the cylinder z is vanishingly small. Appling Gauss‘s law to B field, we have

 B  ds  (B n1  a z  B n 2  a z )dS  (B n1  Bn2 )dS  0


S

or B n1  Bn2  1 H n1   2 H n2 (2.32)

Example 2.36: The magnetic flux density in a finitely conducting cylinder of radius 10
cm and with a relative permeability of 5 is found to vary as 0.2/ a T. If the region
surrounding the cylinder is characterized by free space, determine the magnetic flux
density just outside the cylinder.

Solution: The interface is at =10 cm, the B in the cylinder just beneath the boundary is
given by
0.2 2
Bc  a φ  2a φ T.  H c  a φ  318.31a φ kA/m
0.1 5  4  10 7
As the magnetic flux density is tangential to the boundary and K=0 for the finite
conductivity of the cylinder, from (2.32) we obtain
H a  318.31a φ kA/m
Therefore, the magnetic flux density just above the surface of the cylinder in free space is
B a   0 H a  0.4a T

2.12 ENERGY IN MAGNETIC FIELD


The energy stored in a steady magnetic field is given by
1 1 1 1
Wm  I 2 L  I   IB  dS   I(  A)  dS
2 2 2 2
1 1
Using stock' s theorem we have Wm   IA  dl   A  Jdv
2 2
Now A  J  A  (  H )  H  (  A)    ( A  H )  H  B    ( A  H )
1 1 1 1
Wm 
2  A  Jdv   [H  B    ( A  H)]dv   H  Bdv     ( A  H)dv
2 2 vol 2 vol
1 1
 
2 vol
H  Bdv   ( A  H)  dS
2s
As H varies inversely as square of distance (1/R2), A varies as (1/R) and the surface area
increases as R2, it follows that the second term in Wm is a function of (1/R). In order to
consider all the fields in calculating Wm, we have to consider a very big closed surface
with R, the second term in Wm thus becomes zero.
1 1 1 B2
Wm   H  Bdv   H dv   dv
2
(2.33)
2 vol 2 vol 2 vol 

Example 2.37: Calculate the energy stored in the magnetic field of the toroidal winding
discussed in Example 2.15.

Solution: H inside the toroid is


NI
H a φ for a    b as obtained in Example 2.15. Now from (2.33) we obtain
2
2 2 2
 NI 
2 b
1  NI 
b h h
1 1 1
 Wm   H 2 dv        dddz        dddz
2 vol 2   a  0 z 0  2  2  2   
a 0 z 0

N 2 I 2 h lnb / a 
1

4

2.13 CALCULATION OF INDUCTANCE


The self inductance of a current carrying loop is defined as the magnetic flux linkage per
unit current in the loop itself, i.e., L=/I. The self inductance can also be calculated from
Wm=I2L/2.
The mutual inductance between two circuits is the magnetic flux linkage with one circuit
per unit current in the other. If I1 current in circuit 1 makes a flux linkage 12 in circuit 2,
the mutual inductance L12 between circuit 1 and 2 is given by L12= 12/I1. If B1 be the
flux density due to current I1, the flux linkage of circuit 2 is given by
N 2  B 1  ds 2 N 2  (  A 1 )  ds 2
 12  N 2  B 1  ds 2 . Therefore, L12  
s2 s2

s2
I1 I1
N 2  A 1  dl 2
Using Stock' s theorem, we have L12 
l2

I1
μN1 I 1 dl 1
4 l1 R
Now from (2.12) we have A1 

where N1 and N2 are the no. of turns in circuit 1 and 2, and R is the distance between the
incremental segments dl1 and dl2.

N 2  A 1  dl 2
N 1 N 2 dl 1  dl 2
Therefore L12   
l2
(2.34)
I1 4 l1 l2
R

Equation (2.28) is known as Neumann formula.


Comparing equations (2.30) and (2.34), we may prove that the magnitude of the force
between two conductors can be written as
dL12
F  I1 I 2 (2.35)
dR

Example 2.38: Find the self inductance of some typical geometry without considering
internal inductance.

Solution:
1) Self inductance of a long solenoid
In case of long solenoid, the magnitude of B field in the solenoid is B  IN . If the
cross-sectional area of the solenoid is A, then the flux linkage of the coil is
=N=NBA=IN2A. Therefore the self inductance of the solenoid L=/I=N2A Henry.

2) Self inductance of a Toroid of circular cross-section

For a circular path of radius =R, the ampere‘s circuital law gives us
NI
 H  dl  NI  B  2 R a
If the cross-sectional area of the toroid is A, the flux linkage of the coil is =N=NBA=
IN2A/2R Therefore the self inductance of the toroid L=/I=N2A/2R Henry.

3) Self inductance of a Coaxial cable


d
Assume the current through the inner dz
conductor is I. Then the magnetic field
at a position  meter away from the
centre of inner conductor is
I
 H  dl  I  B  2  a
The magnetic flux that thus produced
I
crosses the area drdz, is d  ddz
2 
The flux linkage per unit length of the cable is
1 D
I
  
z 0 d
2 
ddz.

I 
 ln( D / d ) and the inductance per unit length is L  ln(D / d ) Henry/m
2 2

4) Per unit length inductance of two-wire transmission line


Assume the transmission line is along the z-direction and one at (x=0, y=0) carries a
current of I amp and the other at (x=0, y=R) carries –I amp. The radius of both
conductors is assumed to be d. In this case, the magnetic field B and hence flux crosses
the yz-plane in between the conductors will be additive and is directed along –x axis.
I I
B  [  ]a x . The flux linkage per unit length of the line is
2y 2 ( R  y )
1 Rd
I I I I
   [ 2y  2 ( R  y) ]dydz  2 ln{(R  d ) / d 2}  ln{( R  d ) / d }.
2

z 0 y  d

R

dy

dz
Therefore the inductance for unit length of the two wire transmission line is L=(/)
ln{(R-d)/d} (/) ln(R/d) H/m.

Example 2.39: Find the magnetic field at radius r within a solid copper conductor of
radius ro>r carrying current I uniformly distributed over the cross-section. Also determine
the internal inductance of the conductor.

Solution:The current that crosses an area of r2=I r2/ r02. Therefore on applying the
ampere‘s circuital law along the closed circular path of radius r is
r 2  r 2 r
 H  dl  ( r0 ) I  B  2r ( r0 ) Ia  2r02 Ia
Now the storage magnetic field is obtained as
r 2 z 2
1 B2 1 0 1  r  I 2 
Wm   dv   0 0   2r02 I  rdrddz  zJ L H /m
2v  20  16 8

Example 2.40: Calculate the internal inductance of a thick conducting cylinder of inner
radius a and outer radius b, as shown in the following figure.

Solution:

Assume that a current I be flowing through the cylindrical tube and having a uniform
current density throughout the cross-section. The current density J is then
I
J az .
 (b  a 2 )
2

From ACL, we obtain  H  dl  I   J  ds


l s

H will be along  direction; if a circular path of radius  (a    b), we obtain


 2
I I ( 2  a 2 )
H  2     (b 2  a 2 ) z
a  dda z 
a 0 (b 2  a 2 )
I ( 2  a 2 ) I ( 2  a 2 )
H  H aφ
2 (b 2  a 2 ) 2 (b 2  a 2 )
2 2 b 2 z 2
1  I ( 2  a 2 )  1  I   ( 2  a 2 ) 
Wm     2 
dv    2        dddz
2 v  2 (b  a ) 
2
2  2 (b  a ) 
2
a 00
2 b 2 z
1  I   3 a4 
        2 a    dddz
2

2  2 (b 2  a 2 )  a 00


 z 
I
 
2
  b4  a4

a 2 2
b  a 2
 a
4
ln
b
  
 2 (b  a )  
2 2
4 a

L 



 b4  a4
 a 2 2
b  a

2
 a 4
lnb
  H /m
2 (b 2  a 2 ) 2  4 a

Example 2.40: Show that the mutual inductance between a straight long conductor and a
coplanar equilateral triangular loop shown in the following figure is
0 
 a  b ln a  b  a 
3  b 

Solution:

600

x
r

600
b a

I
B at a distance r from the straight long conductor is B  aφ
2 r
a b
0 I
Therefore,    B  ds   xdr
s b
2 r

However, (x/2)=(a+b-r)tan 300  x=(2/3){a-(r-b)}


a b a b
0 I 1 0 I  a  b   I  ab 
 
2
 a  (r  b)dr     1dr  0 (a  b) ln  a
3 b 2 r 3  b  r  3  b 
 0  ab 
M   (a  b) ln  a
I 3  b 

2.14 APPLICATION OF MAGNETOSTATIC FIELD


In the following, we will introduce few examples where the magnetostatic field is widely
used.

Magnetic separator
An important application of magnetostatic field is found in a device called a magnetic
separator, which is designed to separate magnetic materials from nonmagnetic ones. The
device is shown in the following figure.
A mix of magnetic and nonmagnetic materials is fed on an endless belt running at a
constant speed. The belt passes over a magnetic pulley which consists of an iron shell
containing an exciting coil that produces the magnetic field. The nonmagnetic material
immediately drops off into a bin while the magnetic material is held by the pulley until
the belt leaves the pulley. The magnetic material is therefore carried further round the
pulley and then dropped into a second bin as shown in the figure.

Cyclotron
Cyclotron is used to generate beams of high energy charged particles, such as protons or
deuterons, required to investigate the subatomic structure of an atom.
For this purpose, a cyclotron requires one electron gun through which the charged
particle is made to pass again and gain. In its simplest form, a cyclotron consists of two
D-shaped cavities made of copper, as shown in the following figure.
A high frequency oscillator is across the two cavities. As expected, the electric field will
exist only within the gap between the cavities, and the charged particle will gain energy
only while passing through the gap. The two cavities are sealed in a vacuum chamber to
minimize the loss in energy due to collisions with air molecules. The whole structure is
immersed in a uniform magnetic field.
The action begins when the charged particle is accelerated by the electric field in the gap
and enters one of the two D-shaped cavities. Once the charged particle is inside the
cavity, it follows a semicircular path. As there is no electric field within the cavity, the
velocity of the charged particle remains the same. If the frequency of oscillation is the
same as the cyclotron frequency, the applied voltage will reverse its polarity by the time
the charged particle reaches the gap. The reversal of the applied voltage changes the
direction the direction of the electric field within the gap and accelerates the charged
particle into the other D-shaped cavity, where the particle describes another circular path
of somewhat larger radius. Thus the particle gains KE each time it crosses the gap,
thereby moving into an orbit of larger radius. This process continues until the charged
particle reaches the outer edge of the D-shaped cavity, where it is ejected out.

Maglev Train
A brief review of magnets will help explain how maglev (magnetic levitation) trains
work. Every magnet has a north pole and a south pole. Similar poles of two magnets repel
each other; opposite poles attract each other. These principles govern the levitation of
maglev trains.
Permanent magnets are always magnetic. Electromagnets are magnetic only when an
electric current flows through them. The north and south poles of an electromagnet are
related to the direction of the current. If the direction of the current is reversed, the poles
are reversed.
In maglevs that levitate by magnetic repulsion, the train lies over the guide way. Magnets
on top of the guide way are oriented to repel similar poles of magnets in the bottom of the
maglev. This pushes the train upward into a suspended position. This system is designed
for maglevs that contain groups of extremely powerful superconducting electromagnets.
These magnets use less electricity than conventional electromagnets, but they must be
cooled to very low temperatures—from −269 degrees Celsius to −196 degrees Celsius.
In maglevs that levitate by magnetic attraction, the bottom of the train wraps around the
guide way. Levitation magnets on the underside of the guide way are positioned to attract
the opposite poles of magnets on the wraparound section of the maglev. This raises the
train off the track. The magnets in the guide way attract the wraparound section only
strongly enough to raise the train a few centimeters into a ―floating‖ position. The
wraparound section does not touch the guide way.
To picture how a maglev train is propelled forward, think of three bar magnets lined up
on the floor. The magnet in front is pulling with an attracting (opposite) magnetic pole
and the magnet in back is pushing with a repulsing (similar) magnetic pole. The magnet
in the middle moves forward. A maglev's guide way has a long line of electromagnets.
These pull the train from the front and push it from behind. The electromagnets are
powered by controlled alternating currents, so they can quickly change their pull and
push poles, and thus continually propel the train forward.
SUMMARY
1) Biot-Savart‘s law states that the H-field resulting from a current carrying conductor is
I dl   (r  r )
4  r  r  3
H

2) Ampere‘s circuital law,  H  dl  I   B  dl  μI or   H  J    B  J


can also be used to determine the magnetic field.
3) The Gauss‘s law   B =0 indicates that the magnetic field is solenoidal and there is
magnetic monopole.
4) The vector magnetic potential A is related with B as B    A.
5) The vector magnetic potential A for a current carrying conductor is
μ 1
A(r )  
4 r - r
I(r )dl (r )

6) The magnetic flux can be obtained from    B  ds   (  A )  ds   A  dl


7) The force experienced on a moving charge in static electromagnetic field is
F=Q(E+uB).
8) The force between two current carrying conductors is attractive if the currents flow in
the same direction, and the force is repulsive if the currents flow in opposite
directions.
9) The boundary conditions at the interface of two generalized media are Ht2-Ht1=K and
B n1  Bn2 .
10) K=0 across the boundary of almost all physical media; but K0 only when at least
one of the media is conductor.
11) The energy stored in the magnetic field is
1 1 B2
Wm       dv.
2
H dv
2 vol 2 vol
12) The mutual inductance between two circuits is obtained by Neumann formula
N1N 2 dl1  dl 2
4 l1 l2
L12 
R

SAMPLE QUESTIONS
2.1 Which one of the following is not a source of magnetostatic field?
(a) A dc current flowing through a wire
(b) A permanent magnet
(c) An accelerated charge
(d) A charged disk rotating at uniform speed
2.2 Identify the configuration in the following figure that is not a correct
representation of I and H.

2.3 Plane y=0 carries a uniform current of 30 az mA/m. At (1, 10, -2), the
magnetic field intensity is
(a) -15 ax mA/m (b) 15 ax mA/m (c) 477.5 ay A/m (d) 18 ay nA/m
(e) None of the above
2.4 For the currents and closed paths of the following figure, calculate the
value of  H  dl .
2.5 Explain Biot-Savart‘s law.
2.6 State and explain Ampere‘s circuital law.
2.7 Derive the equation of vector magnetic potential for a current carrying
conductor.
2.8 Explain Gauss‘s law for magnetic field.
2.9 What is the physical meaning of   B =0?
2.10 Explain Lorentz force law.
2.11 Show that the magnetic force on a charge do no work.
2.12 Make a comparison between electric and magnetic forces on a
charged particle.
2.13 Show that the force between two current carrying conductors is
attractive if the currents flow in the same direction, but it is repulsive
if the currents flow in opposite directions.
2.14 Determine the boundary conditions for magnetostatic field at the
interface between generalized media.
2.15 Express the energy stored in magnetic field in terms of magnetic field
intensity/magnetic field density.
2.16 Derive the Neumann formula.

You might also like