Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 31

Chapter 3

Magnetostatic Fields (charges moving with constant velocity=direct current)


⤿ In chapter 2, we have seen static electric fields characterized by E or D.
⤿ In chapter 3, we will see static magnetic fields, which are characterized by H or B.
⤿ There are similarities and dissimilarities between electric and magnetic fields.
⤿ As E and Dare related according to D=ε E for linear material space, H and B are related
according to B=μ H .
⤿ The analogy is presented here to show that most of the equations we have derived for the
electric fields may be readily used to obtain corresponding equations for magnetic fields if
the equivalent analogous quantities are substituted.

An electrostatic field is produced by static or stationary charges.


But a static magnetic (or magnetostatic) field is produced by
 Charges are moving with constant velocity.

1|Page
 A constant current flow (or direct current). This current flow may be due to magnetization
currents as in permanent magnets, electron-beam currents as in vacuum tubes, or
conduction currents as in current-carrying wires.
Applications of magnetostatics:
 motors,  bell ringers,
 transformers,  advertising displays,
 microphones,  memory stores,
 compasses,  magnetic separators,
 telephone  television focusing controls,
 magnetically levitated high speed vehicles, and so on
The two fundamental governing laws for magnetostatic fields are:
(1) Biot-Savart's law (general law)
(2) Ampere's circuit law (special law of Biot-Savart’s law)
NB:
⤿ Like Coulomb's law, Biot-Savart's law is the general law of magnetostatics.
⤿ Just as Gauss's law is a special case of Coulomb's law, Ampere's law is a special case of Biot-
Savart's law and is easily applied in problems involving symmetrical current distribution .

3.1 Biot-Savart's Law


Biot-Savart's law states that the magnetic field intensity dH is
 Directly proportional to the product Idl and the sine of the angle α
 Inversely proportional to the square of the distance R between P and the element.
Where
Idl is differential current element.
α is angle between the current element and the line joining P to the element .

Figure 3.1 magnetic field dH at P due to current element Idl.

Idlsinα kIdlsinα
That is dH ∝ 2
⟹ dH= 2
R R
1
Where k is constant of proportionality and is given by k = 4 π .
Thus,
Idlsinα
dH =
4 πR2

From the definition of cross product, it is possible to put the equation above in vector form as
2|Page
Id l x aR Id l x R
d H= 2
= 3
4 πR 4 πR
where R = |R| and aR = R/R.
Thus, the direction of dH can be determined by the right hand rule as
 if the right-hand thumb shows in the direction of the current,
 The right-hand fingers encircling the wire shows the direction of dH as in Figure 3.2(a).
 Alternatively, by the right-handed screw rule: with the screw placed along the wire and
pointed in the direction of current flow, the direction of advance of the screw is the
direction of dH as in Figure 3.2(b).

Figure 3.2 Determining the direction of dH using


(a) The right-hand rule, or (b) the right-handed screw rule.
Direction Conventions:
 H (or current I) into a page by a small circle with a cross sign
 H (or current I) out of a page by a small circle with a dot sign as in Figure 3.3.

Figure 3.3 Conventional representation of H (or I)


(a) out of the page and (b) into the page.
Just as we can have different charge configurations, we can have different current distributions:
⤿ Line current I
⤿ Surface current density K (in amperes/meter)
⤿ Volume current density J (in amperes/meter square) as shown in Figure 3.4.
⤿ The source elements are related as
Id l≡ K dS ≡ J dv

Thus, Biot-Savart law in terms of the distributed current sources is


Id l x aR
H=∫ 2
(line current )
4 πR
K dSx a R
H=∬ (surface current )
4 πR 2
3|Page
J dvx aR
H=∭ 2
( volume current )
4 πR

Figure 3.4 Current distributions:


(a). line current, (b) surface current, (c) volume current.

Example:
 Apply equation of the line current to determine the field due to a straight current carrying
filamentary conductor of finite length AB as in Figure 3.5.
 Assume the conductor is along the z-axis with its upper and lower ends respectively
subtending angles α 2 and α 1 at P, the point at which H is to be determined.
 The contribution dH at P due to an element dl at (0, 0, z) will be
Id l x R
d H=
4 πR
3 but d l=dz a z∧R=ρ a ρ−z a z so
Iρdz
H=∫ a∅
d l x R=ρdz a∅ Hence, 2
3
2 2
4 π (ρ + z )
From figure 3.5, z=ρcotα , dz=−ρ cosec αdα
2

α2 2 2 α2
−1 ρ cosec αdα −I
H= ∫
4 π α ρ cosec α
3 3
a ∅= a ∫ sinαdα
4 πρ ∅ α
1 1

I
H= (cos α 2−cos α 1 ) a∅
4 πρ

4|Page
Figure 3.5 Field at point P due to a straight filamentary conductor.

Notice:
 This formula is general for any straight filamentary conductor of finite length.
 H is always along the unit vector a ∅ (i.e., along concentric circular paths) irrespective of the
length of the wire or the point of interest P.
 Always α 1∧α 2 are measured from line towards R1 and R2 in CCW direction.
 Always α 1 is measured @ the tail and α 2 @ the head of the current.
 ρ is the shortest perpendicular distance from the line to P.
 As a special case, when the conductor is semiinfinite (with respect to P) so that point A is
now at (0, 0, 0) while B is at ( 0,0 , ∞ ) ; α 1=90 0 , α 2=00 , the above equation becomes
I
H= a
4 πρ ∅

 Another special case is when the conductor is infinite in length. For this case, point A is at
( 0,0 ,−∞ ), while B is at( 0,0 , ∞ ) ;α 1=180 0 , α 2=0 0, then H reduces to
I
H= a
2 πρ ∅
 To find unit vector a ∅ in the above equations is not always easy. A simple approach is to
determine a ∅ from
a ∅=al x a ρ
where
a l is a unit vector along the line current and
a ρ is a unit vector along the perpendicular line from the line current to the field point.

Example3.1:
The conducting triangular loop in Figure 3.6(a) carries a current of 10 A. Find H at (0, 0, 5) due to
side 1 of the loop.

5|Page
Solution:
The key point in figuring out α 1 , α 2 , ρ∧a∅ .
To find H at (0, 0, 5) due to side 1 of the loop in Figure 3.6(a), consider Figure 3.6(b), where side 1
is treated as a straight conductor.
Notice that we join the point of interest (0, 0, 5) to the beginning (tail) and end (head) of the line
current. Observe that α 1 , α 2∧ρ are assigned in the same manner as in Figure 3.5.

Figure 3.6 (a) conducting triangular loop, (b) side 1 of the loop.

EXERCISE 3.1
Find H at (0. 0, 5) due to side 3 of the triangular loop in Figure 3.6(a).
Answer: -30.63ax + 30.63ay mA/m.

Example 3.2: A circular loop located on x2 + y2 = 9, z = 0 carries a direct current of 10 A along a ∅.


Determine H at (0, 0, 4) and (0, 0, -4).

6|Page
Figure 3.7 (a) circular current loop, (b) flux lines due to the current loop.
Solution:
Consider the circular loop shown in Figure 3.7(a). The magnetic field intensity dH at point
P (0, 0, h) contributed by current element Idl is given by Biot-Savart's law:
Id l x R
d H=
4 πR3
Where d l=ρd ∅ a ∅ , R=( 0,0 , h )− ( x , y ,0 )=−ρ a ρ +h a z∧¿

| |
aρ a∅ a z
d l x R= 0 ρd ∅ 0 =ρhd ∅ a ρ+ ρ2 d ∅ a z
−ρ 0 h
I
3 (
ρhd ∅ aρ + ρ d ∅ az ) =dH ρ a ρ+ dH z a z
2
d H=
Hence,
4 π ( ρ2 + z 2 ) 2

Notice:
⤿ By symmetry, the contributions along a ρ add up to zero because the radial components
produced by pairs of current element 180° apart cancel.
⤿ To show mathematically, since a ρ=cos ∅ ax +sin ∅ a y , integrating cos ∅ or sin ∅
Over 0 ≤ ∅ <2 π gives zero, thereby showing that H ρ=0.
⤿ Thus,

I ρ2 d ∅ a z I ρ2 2 π a z I ρ2 a z
H=∫ dH z a z =∫ 3
= 3
= 3
0
4π ( ρ +z ) 4 π (ρ +z ) 2(ρ +z )
2 2 2 2 2 2 2 2 2

⤿ This is the general formula for any circular loop.


(a)Substituting I =10 A , ρ=3 , h=4 gives
H ( 0,0,4 ) =10 ¿¿
(b)The same as of case (a) except h is replaced by –h. Hence,
H ( 0,0 ,−4 ) =H ( 0,0,4 )=0.36 az A /m

7|Page
3.2 Ampere's Circuit Law—Maxwell's Equation
Ampere's circuit law states that
 The line integral of the tangential component of H around a closed path is the same as
the net current I enc enclosed by the path. i.e, the circulation of H equals I enc; that is,
∮ H . d l =I enc
 Is similar to Gauss's law and is used to find H when the current distribution is
symmetrical.
 By applying Stoke's theorem to the left-hand side of the above equation, we obtain
I enc =∮ H . d l=∬ ( ∇ x H ) . d S
Since
I enc =∬ J . d S

 Comparing these two double(surface) integrals, we obtain


∇ x H=J
 This is called the third Maxwell’s equation; it is Ampere's law in differential form.
 We observe that ∇ x H=J ≠ 0; that is, magnetostatic field is not conservative.

3.3 Applications Of Ampere's Law


⤿ We will apply Ampere's circuit law to determine H for some symmetrical current
distributions as we did for Gauss's law.
⤿ We will consider
 an infinite line current,
 an infinite current sheet, and
 an infinitely long coaxial transmission line .

A. Infinite Line Current


 Consider an infinitely long filamentary current I along the z-axis as in Figure 3.8.
 To determine H at an observation point P, we allow a closed path pass through P. This path,
on which Ampere's law is to be applied, is known as an Amperian path (analogous to the
term Gaussian surface).
 We choose a concentric circle as the Amperian path in view of the above equation, which
shows that H is constant provided ρ is constant.
 Since this path encloses the whole current I, according to Ampere's law
I =∫ H ∅ a ∅ . ρd ∅ a∅ =H ∅∫ ρd ∅=H ∅ .2 πρ or
I
H= a as expected
2 πρ ∅

8|Page
Figure 3.8 Ampere's law applied to an infinite filamentary line current
B. Infinite Sheet of Current
⤿ Consider an infinite current sheet in the z = 0 plane.
⤿ If the sheet has a uniform current density K = Kyay A/m as shown in Figure 3.9, applying
Ampere's law to the rectangular closed path (Amperian path) gives
∮ H . d l =I enc =K y b
⤿ To evaluate the integral, we regard the infinite sheet as comprising of filaments; dH above
or below the sheet due to a pair of filamentary currents.

Figure 3.9 Application of Ampere's law to an infinite sheet:


(a) Closed path 1-2-3-4-1,
(b) Symmetrical pair of current filaments with current along ay.

⤿ As evident in Figure 3.9 (b), the resultant dH has only an x-component. Also, H on one side
of the sheet is the negative of that on the other side. Due to the infinite extent of the sheet,
the sheet can be regarded as consisting of such filamentary pairs so that the characteristics
of H for a pair are the same for the infinite current sheets, that is,

where H 0 is yet to be determined.


⤿ Evaluating the line integral of H along the closed path in Figure 3.9 (a) gives
∮ H . d l =¿
9|Page
1
We obtain H 0= 2 K y. Substituting H 0 gives

{
1
K y a x , z >0
H= 2
−1
K a , z <0
2 y x

⤿ In general, for an infinite sheet of current density K A /m,


1
H= K x an
2

Where a n is a unit normal vector directed from the current sheet to the point of interest.
C. Infinitely Long Coaxial Transmission Line
 Consider an infinitely long transmission line consisting of two concentric cylinders having
their axes along the z-axis.
 The cross section of the line is shown in Figure 3.10, where the z-axis is out of the page.
 The inner conductor has radius a & carries current I while the outer conductor has inner
radius b, thickness t & carries return current - I.
 We want to determine H everywhere assuming that current is uniformly distributed in
both conductors.
 Since the current distribution is symmetrical, we apply Ampere's law along the Amperian
path for each of the four possible regions: 0 ≤ ρ≤ a , a ≤ ρ ≤ b ,b ≤ ρ ≤ b+t∧ρ ≥ b+t .

Figure 3.10 Cross section of the transmission Figure 3.11 Plot of H against p.
line; the positive z-direction is out of the page.
(a) For region 0 ≤ ρ≤ a , we apply Ampere’s law to path L1, giving

∮ H . d l=I enc =∫ J . d S
L1

Since the current is uniformly distributed over the cross section,


I
J= a z , d S=ρd ∅ dρ a z then
π a2
2
I I Iρ
I enc =∫ J . d S= 2 ∬ ρd ∅ dρ= 2 π ρ 2= 2
πa πa a
10 | P a g e
2

Then H ∅ ∫ dl=H ∅ 2 πρ= 2 or
a

H∅= 2
2π a

(b) For region a ≤ ρ ≤b , we apply Ampere’s law to path L2, giving


∮ H . d l=I enc =I ⟹ H ∅ 2 πρ=I ⟹ H ∅ = 2 Iπρ


L2

Since the whole current I is enclosed by L2


(c) For region b ≤ ρ ≤b +t , we apply Ampere’s law to path L3, giving

∮ H . d l=H ∅ 2 πρ=I enc where I enc =I +∫ J . d S


L3

And J in this this case is the current density (current per unit area) of the outer
conductor and is along -az that is,
−I
J= az
π [(b+ t ) −b ]
2 2

[ ]
2π ρ
I ρ2 −b2
Thus, I enc =I − ∫∫ ρdρd ∅=I 1−
π [ ( b+ t ) −b 2 ] ∅ =0 ρ=b t 2+ 2bt
2

Substituting this
I
2 πρ
H∅=
1− 2
[
ρ2 −b2
t + 2bt ]
(d). For region ρ ≥ b+t , we apply Ampere’s law to path L4 , giving

∮ H . d l=I −I =0 ⟹ H ∅=0
L4

⤿ Putting all the above together;

⤿ The magnitude of H is sketched in Figure 3.11.


⤿ Ampere's law can only be used to find H due to symmetric current distributions for which
it is possible to find a closed path over which H is constant in magnitude.
Example: Planes 2 = 0 and z = 4 carry current K = -10ax A/m and K = 10ax A/m, respectively.
Determine H at (a) (1, 1, 1) (b) (0, - 3, 10)

11 | P a g e
Figure 3.12 parallel infinite current sheets.
Solution:
Let the parallel current sheets be as in Figure 3.12. Also let H=H 0+ H 4, where H 0 and H 4 are the
contributions due to the current sheets z = 0 and z = 4, respectively.

3.4 Magnetic Flux Density-Maxwell's Equation


 The magnetic flux density B is similar to the electric flux density D.
 As D=ε 0 E in free space, the magnetic flux density B is related to the magnetic field intensity
H according to
B=μ 0 H
where μ0 is known as the permeability of free space. It is in henrys/meter (H/m) and has the
value of
−7
μ0 =4 πx 10 H / m

 The magnetic fluxΨ through a surface S is given by



Ψ =∬ B .d S
S

where Ψ is in Webers (Wb) and the B is in webers/square meter (Wb/m2) or teslas.


 The magnetic flux line is the path to which B is tangential at every point in a magnetic field.
 It is the line along which the needle of a magnetic compass will orient itself if placed
in the magnetic field.
 For example, the magnetic flux lines due to a straight long wire are shown in Figure 3.13.
The flux lines are determined using the same principle followed in for the electric flux lines.
 The direction of B is taken as that indicated as "north" by the needle of the magnetic
compass. Notice that each flux line is closed and has no beginning or end.
 It is generally true that magnetic flux lines are closed and do not cross each other
regardless of the current distribution.
12 | P a g e
Figure 3.13 Magnetic flux lines due to a straight wire with current coming out of the page.

 In an electrostatic field, Ψ =∯ D . d S=Q .


 Thus, it is possible to have an isolated electric charge as shown in Figure 4.14(a), which also
reveals that electric flux lines are not necessarily closed.
 Unlike electric flux lines, magnetic flux lines always close upon themselves as in Figure
3.14(b). It is b/c, it is not possible to have isolated magnetic poles (or magnetic charges).
 For example, if we desire to have an isolated magnetic pole by dividing a magnetic bar
successively into two, we end up with pieces each having north and south poles as illustrated
in Figure 3.15, we find it impossible to separate the north pole from the south pole.

Figure 7.17 Flux leaving a closed surface due to:


(a) Isolated electric charge, Ψ =∯ D . d S=Q
(b) Magnetic charge, Ψ =∯ D . d S=0

 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 . d S=0
 This equation is referred to as the law of conservation of magnetic flux or Gauss's
law for magnetostatic fields just as Ψ =∯ D . d S=Q is Gauss's law for electrostatic
fields.
 Although the magnetostatic field is not conservative, magnetic flux is conserved.
By applying the divergence theorem to the above equation, we obtain

13 | P a g e
∯ B . d S=0=∭ (∇ . B) dv=0 ⟹ ∇ . B=0
 This equation is the fourth Maxwell's equation. This Equation shows as
⤿ Magnetostatic fields have no sources or sinks.
⤿ Magnetic field lines are always continuous.

Figure 3.15 Successive division of a bar magnet results in pieces with north and south poles,
showing that magnetic poles cannot be isolated.
3.5 Maxwell's Equations For Static EM Fields
TABLE: Maxwell's Equations for Static EM Fields
Differential Form Integral Form Remarks
∇ . D=ρv ∯ D. d S=∭ ρ v dv Gauss’s law
∇ . B=0 ∯ B . d S=0 Nonexistence of magnetic monopole
∇ x E=0 ∮ E . d l=0 Conservativeness of electrostatic field
∇ x H=J ∮ H . d l =∬ J . d S Ampere's law
NB:
 The choice between differential and integral form depends on a given problem.
 As of the table a vector field is defined completely by specifying its curl and divergence.
 A field can only be electric or magnetic if it satisfies the corresponding Maxwell's equations.
 Maxwell's equations as in the above table are only for static EM fields.
 The divergence equations will remain the same for time-varying EM fields but the curl
equations will have to be modified.

3.6 Magnetic Scalar & Vector Potentials


 We recall that some electrostatic field problems were simplified by relating the electric
potential V to the electric field intensity E (E=−∇ V ).
 Similarly, we can define a potential associated with magnetostatic field B.
 In fact, the magnetic potential could be scalar V m or vector A. To define V m and A involves
recalling two important identities

∇ x ( ∇ V )=0
∇ . ( ∇ x A ) =0

14 | P a g e
which must always hold for any scalar field V and vector field A.
 Just as E=−∇ V , we define the magnetic scalar potential V m (in amperes) as related to H
according to
H=−∇ V m if J=0
J=∇ x H =∇ x (−∇ V m )=0
since V m must satisfy the condition in the above equation.
 Thus, the magnetic scalar potential V m is only defined in a region where J=0.
 Also V m satisfies Laplace's equation just as V does for electrostatic fields; hence,

 We know that for a magnetostatic field,∇ . B=0, as stated earlier.


 In order B to satisfy conditions ∇ . B=0 and ∇ . ( ∇ x B )=0 , we can define the vector magnetic
potential A (in Wb/m) such that
B=∇ x A
dQ
 Just as we defined V =∫ 4 π ε r we can define
0
μ0 Id l
A=∫ (for line current )
4πR
μ 0 K dS
A=∬ (for surface current )
4π R
μ0 J dv
A=∭ ( for volume current )
4π R

 By applying Stoke’s theorem


❑ ❑ ❑ ❑
Ψ =∬ B .d S=∬ (∇ x A) . d S=∮ A . d l⟹ Ψ =∮ A . d l
S S L L

 Thus, the magnetic flux through a given area can be found using either from B or A.
 Also, the magnetic field can be determined using either V m or A; the choice is dictated by
the nature of the given problem except that V m can only be used in a source-free region.
 The use of the A provides a powerful, elegant approach to solving EM problems,
particularly those relating to antennas. It is more convenient to find B by first finding A in
antenna problems.

Example:
2
−ρ
Given the magnetic vector potential A= a z Wb /m, calculate the total magnetic flux
4
π
crossing the surface ∅= 2 , 1 ≤ ρ ≤ 2 m, 0 ≤ z ≤5 m

Solution:
We can solve this problem in two different ways:
−∂ A z ρ
Method 1: B=∇ x A= a∅ = a ∅ ,∧d S=dρdz a∅
∂ρ 2

15 | P a g e
❑ 5 2
1 1 2 1
Hence, Ψ =∬ B .d S= ∫ ∫ ρdρdz= ρ ¿2 (5 )=3.75 Wb
S 2 z=0 ρ=1 4

Method 2: We use Ψ =∮ A . d l=Ψ 1 +Ψ 2 +Ψ 3 +Ψ 4


L

Where L is the path bonding surface S;Ψ 1 ,Ψ 2 , Ψ 3∧Ψ 4 are respectively, the evaluations of ∫ A . d l
along the segments of L labeled 1 to 4. Since A has only a z-component, Ψ 1=0=Ψ 3 .

[ ]
5 0
−1 −1
That is Ψ =Ψ 2+ Ψ 4 = ( 1 )2∫ dz + ( 2 )2∫ dz = ( 1−4 )( 5 ) =3.75Wb
4 0 5 4

EXERCISE
A current distribution gives rise to the vector magnetic potential
A=x y a x + y x a y −4 xyz az Wb /m. Calculate
2 2

(a) B at (-1 , 2 , 5)
(b) The flux through the surface defined by z=1 , 0 ≤ x ≤1 ,−1 ≤ y ≤ 4
Wb
Answer: ( a ) 20 a x + 40 a y +3 az , ( b ) 20 Wb
m2

3.7 Magnetic Forces, Materials & Devices


 In this topic, we will study the force a magnetic field exerts on
 Charged particles,
 Current elements, and
 Loops.
 Such a study is important to problems on electrical devices such as

 Ammeters,  galvanometers,
 Voltmeters,  cyclotrons,
16 | P a g e
 plasmas,  magnetohydrodynamic
 motors, and generators.

3.7.1 Forces Due To Magnetic Fields


Three ways in which force due to magnetic fields can be experienced. The force can be
(a) Due to a moving charged particle Q in an external B field,
(b) On a current element in an external B field, or
(c) Between two current elements.

A. Force on a Charged Particle Q


 Recall that the electric force F eon a stationary or moving electric charge Q in an external
electric field E is given by Coulomb's experimental law as
F e =Q E

 This shows that if Q is positive, F e and E have the same direction.


 A magnetic field B can exert force only on a moving charge. From experiments, it is found
that the magnetic force F m experienced by a charge Q moving with a velocity u in a
magnetic field B is given by
F m=Q u x B
 This clearly shows that F m is perpendicular to both u and B.
 A comparison between F e and the magnetic force F m can be made.
 F e is independent of u and can perform work on the charge and changes its kinetic
energy.
 Unlike F e, F m depends on u and is normal to it. F m cannot perform work because it is
at right angles to the direction of motion of the charge F m . d l=0; it does not cause an
increase in kinetic energy of the charge.
 The magnitude of F m is generally small compared to F e except at high velocities.
 For a moving charge Q in the presence of both electric and magnetic fields, the total force
on the charge is given by
F=F e + F m=Q( E+ u x B)
This is known as the Lorentz force equation. It relates mechanical force to electrical force.
 If the mass of the charged particle moving in E and B fields is m, by Newton's second law of
motion
du
F=m =Q ( E+u x B )
dt
 The solution to this equation is important in determining the motion of charged particles in
E and B fields. But energy transfer can be only by means of the electric field.

17 | P a g e
B. Force on a Current Element
 To determine the force on a current element Id l of a current-carrying conductor due to the
magnetic field B, using the fact that for convection current J= ρv u .
 We recall the relationship between current elements: Id l≡ K dS ≡ J dv . Combining yields
Id l=ρv u dv=dQ u
Alternatively,
dQ dl
Id l= d l=dQ =dQ u
dt dt
Hence,
Id l=dQ u
 This shows that an elemental charge dQ moving with velocity u (thereby producing
convection current element dQ u) is equivalent to a conduction current element Id l.
 Thus, the force on a current element Id l in a magnetic field B is found by merely replacing
dQ u by Id l.; that is,
d F=Id l x B

 If the current I is through a closed path L or circuit, the force on the circuit is given by

F=∮ Id l x B
L

 The magnetic field produced by the current element Id l does not exert force on the element
itself just as a point charge does not exert force on itself.
 The B field that exerts force on Id l must be due to another element. In other words, the B
field in the above equations is external to the current element Id l.
 If instead of the line current element Id l, we have surface current elements K d S or a
volume current element J dv , the magnetic forces will be respectively,

 The magnetic field B is defined as the force per unit current element.
Fm
 Alternatively, B may be defined as the vector which satisfies =u x B just as we defined
Q
Fe
electric field E as the force per unit charge, . Both of these definitions of B show that B
Q
describes the force properties of a magnetic field.

C. Force between Two Current Elements


 Consider the force between two elements I 1 d l1 and I 2 d l 2. According to Biot-Savart's law,
both current elements produce magnetic fields.
 So, we may find the force d ( d F 1) on element I 1 d l1 due to the field d B 2 produced by
element I 2 d l 2 as shown in Figure 3.18.

18 | P a g e
Figure 3.18 Force between two current loops
 This equation is essentially the law of force between two current elements and is
analogous to Coulomb's law, which expresses the force between two stationary charges.
 From the above equation, we obtain the total force F, on current loop 1 due to current loop
2 shown in Figure 3.18 as
μ 0 I 1 I 2 ❑ ❑ d l 1 x (d l2 x aR )
F 1=
4π L L
∮∮ 2
R 21
21

1 2

 The force F 2 on loop 2 due to the magnetic field B1 from loop 1 is obtained by interchanging
subscripts 1 and 2. It can be shown that F 2=−F1 . Thus, F 1 and F 2 obey Newton's third law
that action and reaction are equal and opposite.
Example: A charged particle of mass 2 kg and charge 3 C starts at point (1, - 2 , 0) with velocity
4ax + 3az m/s in an electric field 12ax + l0ay, V/m. At time t = 1 s, determine
(a) The acceleration of the particle (b) Its velocity
(c) Its kinetic energy (d) Its position

Solution:
This is an initial-value problem because initial values are given. According to Newton's second
law of motion, F = ma = QE
where a is the acceleration of the particle. Hence,

Equating components gives


19 | P a g e
Example: A charged particle moves with a uniform velocity 4ax m/s in a region where E = 20ay
V/m and B = Boaz Wb/m2. Determine Bo such that the velocity of the particle remains constant.

20 | P a g e
Figure 3.19 A velocity filter for charged particles
Solution:
If the particle moves with a constant velocity, it implies that its acceleration is zero. In other
words, the particle experiences no net force. Hence,

This example illustrates an important principle employed in a velocity filter shown in Figure 3.19.
In this application, E, B, and u are mutually perpendicular so that Q u x B is directed opposite to,
Q E regardless of the sign of the charge. When the magnitudes of the two vectors are equal,
E
QuB=QE ⟹u=
B
This is the required (critical) speed to balance out the two parts of the Lorentz force. Particles
with this speed are undeflected by the fields; they are "filtered" through the aperture. Particles
with other speeds are deflected down or up, depending on whether their speeds are greater or
less than this critical speed.
Example: A rectangular loop carrying current I 2 is placed parallel to an infinitely long
filamentary wire carrying current I 1 as shown in Figure 3.20(a). Show that the force experienced
by the loop is given by
F=
2π [
−μ 0 I 1 I 2 b 1

1
] a N
ρ0 ρ0 +a ρ

21 | P a g e
Figure 3.20 (a) rectangular loop inside the field produced by an infinitely long wire,
(b) forces acting on the loop and wire.
Solution: Let the force on the loop be
F l=F1 + F 2+ F 3 + F 4=I 2∮ d l 2 x B 1
where F 1 , F2 , F3∧F 4 are, respectively, the forces exerted on sides of the loop labeled 1, 2, 3, and 4
in Figure 3.20(b).
Due to the infinitely long wire
μ0 I 1
B 1= a
2 π ρ0 ∅
Hence,

F 1 is attractive because it is directed toward the long wire; that is, F 1, is along −a ρ due to the fact
that loop side 1 and the long wire carry currents along the same direction. Similarly,

The total force F l on the loop is the sum of F 1 , F2 , F3∧F 4 ; that is,
F=
2π [
−μ 0 I 1 I 2 b 1

1
] a N
ρ0 ρ0 +a ρ

22 | P a g e
which is an attractive force trying to draw the loop toward the wire. The force F w on the wire, by
Newton's third law, is F las shown in Figure 3.20 (b).

3.7.2 Magnetization In Materials Classification Of Magnetic Materials


 As dielectric materials are polarized under the influence of an applied electric field, certain
materials can be magnetized under the influence of applied magnetic field.
 Without an external B field applied to the material, the sum of m's (magnetic moments) is
zero due to random orientation as in Figure 3.21(a).
 When an external B field 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
Figure 3.21(b).

Figure 3.21 Magnetic dipole moment in a volume ∆ v :


(a) Before B is applied, (b) After B is applied.

 If most of the magnetic moments stay aligned after the applied magnetic field is removed, a
permanent magnet is formed.

23 | P a g e
 The magnetic susceptibility ( χ m ) and relative permeability(μ¿¿ r ) ¿ are a measure of how
much magnetization occurs in the material.

Figure 3.22 Classification of magnetic materials

Figure 3.23 Typical magnetization (B-H) curve.

3.8 Inductance, Inductor and Magnetic Energy


 The linkage of an inductor defines the total magnetic flux that links the current.
 If the magnetic flux is produced by the same applied current it is called self-inductance.
 If the magnetic flux produced is by another circuit, then it is called Mutual inductance.
 If the circuit has N identical turns, the flux linkage is given by λ=N Ψ .
 Also, if the medium surrounding the circuit is linear, the flux linkage λ is proportional to
the current I producing it; that is,

24 | P a g e
where L is a constant of proportionality called the inductance of the circuit.
 The inductance L is a property of the physical arrangement of the circuit. A circuit or part
of a circuit that has inductance is called an inductor. We may define inductance L of an
inductor as the ratio of the magnetic flux linkage λ to the current I through the inductor;
that is,

 The magnetic energy (in joules) stored in an inductor is expressed in circuit theory as:

 This energy stored in the magnetic field B of the inductor can be expressed in terms of B or
H as (compare with the energy stored by an electric field,W E)

3.8 Magnetic Boundary Conditions


 The conditions that H (or B) field must satisfy at the boundary b/n two different media.
 We use the following two formulas for the derivation:

Gaus s s law for magnetostatics :∯ B. d S=0
'

S

Amper e ' s circuit Law : ∮ H . d l=0
L

 Consider the boundary between two magnetic media 1 and 2, characterized,


respectively by μ1 and μ2 as in Figure 3.7

(a) (b)
Fig 3.7 Boundary conditions between two magnetic media: (a) for B, (b) for H .
 Applying Gauss law as ∆ h→ 0 gives us
B1 n ∆ S−B2 n ∆ S=0 Thus,

B1 n=B2 n∨μ1 H 1n =μ2 H 2 n … … … … …( v 1)

25 | P a g e
 i.e while the normal component of B is continuous, the normal component of H is
discontinous at the boundary
 Applying Ampere’s law to the closed path abcda of Fig 3.7(b), we obtain

As ∆ h→ 0
B 1 t B2 t
H 1 t −H 2t =K ∨ − =K
μ1 μ2
 This shows that the tangential component of H is also discontinuous.
 In the general case
( H 1 t −H 2 t ) x a n 12=K
 If the boundary is free of current or the media are not conductors (for K=0)
B 1t B2 t
H 1 t =H 2t ∨ = … … ..(v 2 )
μ1 μ 2
 Thus, the tangential component of H is continuous while that of B is discontinuous.
 If the fields make an angle θ with the normal to the interface, combining the two
equations ( v 1) and (v 2) above, we get law of refraction for magnetic flux lines
tanθ1 μ1
=
tanθ2 μ2

Chapter 4
Introduction to Time Varying EM Fields and Maxwell Equations
 In Chapters 2 &3, we have restricted our discussions to static, or time invariant, EM fields.
 In this Chapter, we shall examine situations where electric and magnetic fields are
dynamic, or time varying.
 In static EM fields, electric and magnetic fields are independent of each other whereas in
dynamic EM fields, the two fields are interdependent. i.e,
⤿ A time-varying electric field necessarily involves a corresponding time-varying
magnetic field.
⤿ Time-varying EM fields, represented by E(x, y, z, t) and H(x, y, z, t), are of more
practical value than static EM fields.
⤿ Electrostatic fields are usually produced by static electric charges whereas
magnetostatic fields are due to motion of electric charges with uniform velocity
(direct current) or static magnetic charges (magnetic poles);
⤿ Time-varying fields or waves are usually due to accelerated charges or time-varying
currents. Any pulsating current will produce radiation (time-varying fields).
In summary:
Static charges ------------>>> electrostatic fields
Steady currents ----------->>>magnetostatic fields
Time-varying currents---->>> electromagnetic fields (or waves)

FARADAY'S LAW:

26 | P a g e
 He discovered that the induced emf (in volts), in any closed circuit is equal to the time rate
of change of the magnetic flux linkage by the circuit. I.e
dλ dΨ
V emf = =−N
dt dt

 The negative sign shows that the induced voltage acts in such a way as to oppose the flux
producing it. This is known as Lenz's law and it emphasizes the fact that the direction of
current flow in the circuit is such that the induced magnetic field produced by the induced
current will oppose the original magnetic field.
 Alternatively, Faraday’s law can be stated as whenever there is a relative motion b/n a
conductor and a magnetic field, then an induced electromotive force (emf) is created at the
two ends of the conductor thereby, if the two ends are connected a current will flow.

4.1 TRANSFORMER AND MOTIONAL EMFs


The variation of the magnetic flux can be caused in three of the following ways:
1. By having a stationary loop in a time-varying B field
2. By having a time-varying loop area in a static B field
3. By having a time-varying loop area in a time-varying B field

A. Stationary Loop in Time-Varying B Field (transformer emf)


 This happens when a stationary conducting loop is in an external time varying magnetic B
field. The above Equation becomes
❑ ❑
∂B
V emf =∮ E . d l=−∬ .d S
L S ∂t

 This emf induced by the time-varying current (producing the time-varying B field) in a
stationary loop is often referred to as transformer emf in power analysis since it is due to
transformer action.

Figure 4.1 Induced emf due to a stationary loop in a time varying B field.

 By applying Stokes's theorem to the middle of the emf expression, we obtain the following
27 | P a g e
❑ ❑
∂B −∂ B
∬ ( ∇ x E ) .d S=−∬ ∂t
. d S ⟹ ∇ x E=
∂t
S S

B. Moving Loop in Static B Field (Motional emf)


 When a conducting loop is moving in a static B field, an emf is induced in the loop.
 The force on a charge moving with uniform velocity u in a magnetic field B is
F m=Q u x B
Fm
 We define the motional electric field Em as Em = =u x B
Q
 If we consider a conducting loop, moving with uniform velocity u as consisting of a large
number of free electrons, the emf induced in the loop is
❑ ❑
V emf =∮ Em . d l=∮ (u x B). d l
L L

 This type of emf is called motional emf or flux-cutting emf because it is due to motional
action. It is the kind of emf found in electrical machines such as motors, generators, and
alternators.

Figure 4.2 Induced emf due to a moving loop in a static B field.

 Another example of motional emf is shown in Figure 4.2, where a rod is moving between a
pair of rails. In this example, B and u are perpendicular that all the following holds.

NOTE THE FOLLOWING POINTS:

28 | P a g e
⤿ The integral above is zero along the portion of the loop where u = 0. Thus, dl is taken along
the portion of the loop that is cutting the field (along the rod in Figure 4.2), where u has
nonzero value.

⤿ The direction of the induced current is the same as that of Em ∨u x B .


⤿ The limits of the integral in are selected in the opposite direction to the induced current
thereby satisfying Lenz's law. For Figure 4.2, integration over L is along -ay whereas
induced current flows in the rod along ay.

C. Moving Loop in Time-Varying Field


 This is the general case in which a moving conducting loop is in a time-varying magnetic
field. Both transformer emf and motional emf are present. Combining, gives the total emf as

Example 1: A conducting bar can slide freely over two conducting rails as shown in Figure 4.3.
Calculate the induced voltage in the bar

Figure 4.3

The polarity of the induced voltage (according to Lenz's law) is such that point P on the bar is at
lower potential than Q when B is increasing.

(b)This is the case of motional emf:

29 | P a g e
(d)Both transformer emf and motional emf are present in this case. This problem can be
solved in two ways.

Since the motional emf is negligible compared with the transformer emf. By trigonometric identity

30 | P a g e
4.2 MAXWELL'S EQUATIONS IN FINAL FORMS
The following is Maxwell’s equations for time varying EM fields.
(compare with the table in page 14 which is for static EM fields)

TABLE: Generalized Forms of Maxwell's Equations

Differential Form Integral Form Remarks


∇ . D=ρv ∯ D. d S=∭ ρ v dv Gauss’s law
∇ . B=0 ∯ B . d S=0 Nonexistence of isolated magnetic charge

−∂ B −∂
∇ x E=
∂t
∮ E . d l= ∬B.d S
∂t S Faraday’s Law

∂D ∂D
∇ x H=J +
∂t ∮ H . d l=∬ ( J + ∂ t ). d S Ampere's Circuit law

++++++ END ++++++

31 | P a g e

You might also like