Sie sind auf Seite 1von 110

COURSE DESCRIPTION

Course: Electricity & Magnetism


Lecturer: Peter Török

Aims:
To give students a solid understanding of basic electricity and magnetism

Objectives:
On completion of the course, students will
• understand and be able to use Coulomb's Law
• understand the concept of a potential and be able to relate it to the
electric field through E=-grad(V)
• be able to derive expressions for the electric field components around a
dipole and other simple charge distributions
• appreciate the basic properties of the electric dipole
• understand the concept of electric flux
• be able to derive Gauss's Flux Law and apply it to simple charge
distributions
• understand the concept of capacitance
• be able to derive the capacitance and the associated stored energy of
simple structures both in a vacuum and in the presence of a dielectric
medium
• be able to derive the force between the plates of a simple capacitor
• appreciate the contrasting properties of conductors and insulators
• have a basic understanding of the Drude model of electrical conduction
including the concepts of conductivity and resistivity, and the link with
Ohm's law
• have a basic understanding of the origin of magnetism and properties
of the magnetic field
• know the magnetic equivalent of Gauss's Law and the connection with
magnetic monopoles (which may not exist)
• know the basic Lorentz force law for a charge moving in electric and
magnetic fields, and be able to calculate the motion of a charged
particle under these circumstances
• to understand how the Lorentz force underlies the Hall effect
• know how to extend the Lorentz force law to a current element
• know and be able to apply the Biot-Savart Law to calculate the
magnetic field in the presence of simple current configurations
• understand Ampere's law, and be able to use it to calculate the field in
simple cases
• understand what is meant by a magnetic dipole
• understand electromagnetic induction and Faraday's law
• be able to calculate the emf (electromotive force) induced in a simple
circuit as the magnetic field it changes
• appreciate how electric and magnetic fields are modified in a material
medium
• understand the concepts of polarisation, electric displacement, and
dielectric constant and their relationship
• have experience in diamagnetism, paramagnetism, and ferromagnetic,
and have a simple qualitative understanding of what they are in terms
of magnetic dipoles.
!
1st year Electricity and Magnetism
Course outline

◊ Introduction and history of E&M, Motivation for course

A ELECTROSTATICS
1. Electric field
a. Conservation of charges
b. Coulomb’s law, Coulomb force
c. Electric field, field lines
d. Superposition of charges
e. Dipoles
2. Electric flux
a. Electric flux due to a point charge
b. Gauss’s law
3. Electrostatic potential energy (U) and Electric potential (V)
a. Line integrals and conservative forces
b. Electrostatic potential energy
c. Electric potential
d. Grad V = - E
4. Electric media – conductors and insulators
a. Distribution of charges, Faraday cage
b. Charge density
c. Electric breakdown, lightning, corona discharge
d. Capacitors and capacitance
e. Dielectric materials, Displacement field
• Review of Electrostatics, questions, problems, etc

B CURRENTS AND MAGNETISM


5. Steady (DC) currents
a. Currents and Ohm’s law
6. Magnetism
a. Magnetic fields
b. Lorentz force
c. Force on current element
d. Applications: Motors, mass spectrometers, cyclotrons
7. Quantitative magnetism
a. Biot-Savart law
b. Ampère’s law
c. Hall effect
8. Time varying currents, EM induction
a. Faraday’s law
b. Lenz’s law and eddy currents
c. Inductance, self inductance
d. Superconductors
9. Magnetic materials
a. Magnetic dipole moment
b. Magnetic materials, field strength in the presence of
magnetic materials
• Review of Magnetism, questions, problems, etc
◊ Maxwell’s equations and the wave equation (non-examinable)
◊ Modern applications of Electricity and Magnetism (non-
examinable)

Peter Török
30 January 2012
Chapter 4

Triple Integrals

Triple integrals are the natural extensions of double integrals to three dimensions.
The basic physical motivation of such integrals is the same as for double integ-
rals: determining the amount of a quantity, typically expressed as a density, within
a three-dimensional region necessitates performing a triple integral of the quantity
over that region. Just as for double integrals, there are coordinate systems other
than Cartesian that are convenient for integrating over certain types of regions.
We will discuss the two most common of such coordinate systems, circular cyl-
indrical coordinates and spherical polar coordinates, and show how integrals are
transformed into these coordinate systems.

4.1 Integrals in Cartesian Coordinates


Suppose there is a quantity f that represents the density of a physical quantity,
such as the mass or charge density at every point (x, y, z) in a region of three-
dimensional space. The amount of this quantity within a region V is obtained by
integrating over the ranges of x, y, and z that span the interior of V. Since this
calculation involves three separate integrations, it is called a triple integral, and is
written as
ZZZ
f (x, y, z) dx dy dz . (4.1)
V

Following our discussion of double integrals, there are several points to note about
triple integrals:

1. Once the volume V has been specified, the integral has a unique value.

2. The integrals over x, y, and z can be carried out in any order.

41
42 Triple Integrals

3. If f = 1, the integral yields the volume of the integration region:


ZZZ
dx dy dz = V . (4.2)
V

The evaluation of triple integrals proceeds in direct analogy to the cases de-
scribed in Chapter 3 for double integrals. The following examples illustrate the
di↵erent situations that can arise.

Example. Suppose f = xyz, and 1


y
that V is the volume shown in Fig. 4.1.
We must first determine the ranges of
0
the integration variables. The volume V 1
is a cube in the positive octant of space
with one corner at the origin. The points z
(x, y, z) within the cube have coordinates
within the ranges

0  x  1, 0

x
0  y  1, (4.3) 1

0  z  1. Fig. 4.1: The cubic region for the triple in-


tegral in Eq. (4.4).
The triple integral of f = xyz within the
cube in Fig. 4.1 is therefore calculated as
ZZZ Z 1 Z 1 Z 1
1
f (x, y, z)dx dy dz = x dx y dy z dz = . (4.4)
V 8
|0 {z } |0{z } |0{z }
1 1 1
2 2 2

This type of region, where the ranges of x, y, and z are specified independently,
is the simplest for triple integrals. The most general volume of this type is a
rectangular prism aligned with the coordinate axes, where each side is a rectangle
parallel to one of the coordinate planes. The next two examples have a volumes
which do not satisfy these criteria, with the result that the triple integrals become
iterated integrals.

Example. Suppose that f = xyz, as in the preceding example, and V is the


“wedge” shown in Fig. 4.2. We first determine the ranges of the integration vari-
ables. The wedge is bounded from above by the plane y z = 0, with all other
bounding planes lying parallel to coordinate planes. Thus, the range of x is
Triple Integrals 43

1
0

y
x

1 0

Fig. 4.2: The volume for the triple integral in Eq. (4.9).

0  x  1. (4.5)

The triangular sides of the wedge are parallel to the plane x = 0, so the ranges of
the y and z coordinates cannot be specified independently. Referring to Fig. 3.4,
the ranges of these variables are

0  y  1, 0  z  y. (4.6)

An alternative choice is (cf. Fig. 3.4)

z  y  1, 0  z  1. (4.7)

Using the ranges in Eq. (4.6), the triple integral is


ZZZ Z 1 Z 1 Z y
f (x, y, z)dx dy dz = x dx y dy z dz . (4.8)
V 0 0 0

As was the case for double integrals, this is called an iterated integral because the
upper limit of the z-integral is y, which necessitates evaluating this integral before
the y-integral. The x-integral can be carried out independently of the other two.
Thus, carrying out the required integrations,
Z 1 Z 1 Z y Z Z y
1 1
x dx y dy z dz = y dy z dz
0 0 0 2 0 0
| {z }
1
2
Z 1 ! Z
1 1 2y 1 1 3 1
= y dy z = y dy = . (4.9)
2 0 2 0 4 0 16
| {z } | {z }
1 2 1
2
y 4
44 Triple Integrals

The evaluation of this integral with the ranges in Eq. (4.7) is left as an exercise.

Example. Consider now the integration of f = xyz over the volume in Fig. 4.3.
This region is contained in the positive octant, bounded from below by the x-y
plane and from above by the plane x + y + z = 1.
The ranges of the integration variables are ob- 1
y
tained by first observing that, in the x-y plane,
where z = 0, the (x, y) coordinates within V are 0
bounded by the line x + y = 1. Hence, the ranges 1
of x and y may be chosen as

0  x  1, 0y1 x. (4.10) z

The lower bound for the range of z for all values


0
of x and y is z = 0. The upper bound is obtained
from the equation of the plane, solved for z: z = x
1 x y. Hence, 1

0z1 x y, (4.11) Fig. 4.3: The for the triple integral


in Eq. (4.4).
so the integral to be evaluated is
$ Z 1 Z 1 x Z 1 x y
f (x, y, z)dx dy dz = x dx y dy z dz . (4.12)
V 0 0 0

This is again an iterated integral in which the z-integration must be evaluated first,
then the y-integration, and finally the x-integration. The integral over z is evaluated
as Z 1 x y !
1 21 x y 1
z dz = z = (1 x y)2 (4.13)
0 2 0 2
By substituting this result into the y-integral and carrying out an integration by
parts, we obtain
Z
1 1 x
y(1 x y)2 dy
2 0
Z
1 3
1 x
1 1 x
= y(1 x y) + (1 x y)3 dy
|6 {z 0
} 6 0
0
1 x
1
= (1 x y)4
24 0

1
= (1 x)4 . (4.14)
24
Triple Integrals 45

Finally, substitution of this expression into the x-integral and again integrating by
parts yields
Z 1 1 Z 1
1 4 1 1
x(1 x) dx = x(1 x)5 + (1 x)5 dx
24 0 | 120 {z }0 120 0

0
1
1 1
= (1 x)5 = (4.15)
720 0 720

as the value of the integral in Eq. (4.12).

Example. It is possible also to solve this problem from first principles. Ima-
gine that the volume is divided up into slices parallel to the x y plane. Each slice
in the z direction has a thickness of z as shown in Fig. 4.4. When the volume is
viewed from above (with z axis pointing towards the viewer) the lines that limit
each section look like it is shown on the right hand side in Fig. 4.4. The equation
of the line delimiting the nth slice is y = (1 n z) x.

x y

Fig. 4.4: Figure for the derivation of the volume integral from first principles.

We want to write a volume integral for the entire object and for this we proceed
as follows: first we determine the area of each one of the triangular slices. The
volume of the given slice will be just obtained as the area times the height which
of course is z. Finally we will some together the contribution from all slices and
calculate the limit for z ! 0. This solution therefore exemplifies how a volume
integral problem can be reduced to an area integration problem.
The area of a given triangle is determined by using a double integral. As
Fig. 4.4 shows the limit for x integration will be 0  x  1 n x as a line
intersects the x axis at 1 n z. The corresponding y limit is 0  y  1 n z x.
46 Triple Integrals

Therefore Z Z
1 n x 1 n z x
area = dx dy
0 0
and the volume of an individual sice is therefore given by
Z 1 n x Z 1 n z x
volume = dx dy z
0 0

Because we need to integrate the function f (x, y, z) = xyz we need to weight the
di↵erential volume by this function. Therefore we obtain the value of the integral
for a given slice: Z Z 1 n x 1 n z x
x dx y dy (n z) z.
0 0
The volume integral can now be written as a Riemann sum by:
N (Z
X 1 n z Z 1 n z x ) Z 1 Z 1 z Z 1 z x
lim x dx y dy (n z) z = z dz x dx y dy.
N!1 0 0 0 0 0
n=0

This integral can readily be evaluated in the following way:


Z 1 Z 1 z Z 1 z x Z Z 1 z
1 1
z dz x dx y dy = z dz x(1 z x)2 dx
0 0 0 2 0 0
Z 1
1 1
= z(z 1)3 dz =
24 0 720
Cartesian coordinates are convenient for evaluating triple integrals within volumes
bounded by planes. But there are many situations where other geometries are
used, the most common of which are spheres and volumes contained within sur-
faces of revolution. In the next two sections, we will discuss two coordinate sys-
tems that considerably extend the capabilities of triple integrals.

4.2 Cylindrical Polar Coordinates


4.2.1 Definition of the Coordinate System
Cylindrical polar coordinates generalize circular polar coordinates (Sec. 3.2) to
three dimensions by adding the “height” z to indicate the position of a point rel-
ative to the x-y plane [Fig. 4.5(a)]. The complete transformation from (x, y, z) to
(r, , z) is
x = r cos , y = r sin , z = z, (4.16)
Triple Integrals 47

where
0  r < 1, 0 < 2⇡, 1 < z < 1. (4.17)
This transformation is depicted in Fig. 4.5. The expressions for r and in terms
of x and y are the same as those in Eq. (3.29).

4.2.2 The Integration Element


The integration element of this coordinate system can be obtained in two ways.
The simplest way is to observe that the z-coordinate simply adds a “thickness” dz
to the integration element in circular polar coordinates:

dV = r dr d dz . (4.18)

The other method, described in Problem Set 4, is based on writing any point
(x, y, z) as a radius vector r

r = r cos i + r sin j + z k, (4.19)

and calculating the integration element from the vector product

dV = |drr ·dr ⇥ drz | , (4.20)

where drr , dr , and drz are the di↵erential changes of r with respect to r, , and
z, respectively:

drr = dr cos i + dr sin j, (4.21)

z
y
r

x
(a) (b)
Fig. 4.5: Two illustrations of circular polar coordinates. (a) The definitions of r, , and z. (b)
The representation of any point as the intersection of the surface of constant r (the cylinder),
constant (the vertical plane), and constant z (the horizontal plane).
48 Triple Integrals

dr = r sin d i + r cos d j , (4.22)

drz = dz k . (4.23)

4.2.3 Triple Integrals in Cylindrical Polar Coordinates

z
1
1

0
0 y
x

1 1

Fig. 4.6: The unit upper half-sphere, x2 + y2 + z2 + 1, for z 0.

Example. Consider the sphere with unit radius in the upper half-space, as
shown in Fig. 4.6. The equation of the surface is

x2 + y2 + z2 = 1 , (4.24)

where z 0. To calculate this volume as an integral in circular polar coordinates,


we must first determine the ranges of the integration variables. The ranges of r
and span the interior of the half-sphere:

0  r  1, 0 < 2⇡ (4.25)

The upper bound for the range of z is obtained from the equation of the sphere,
solved for z:
z2 = 1 x 2 y 2 = 1 r 2 . (4.26)
The half-sphere is bounded
p from below by the x-y plane, where z = 0. Hence,
the range of z is 0  z  1 r2 . Thus, the volume integral of the half-sphere is
given by
p
Z 1 Z 2⇡ 1 pZ 1 r2 Z
V = r dr
d dz = 2⇡ r 1 r2 dr
0 0 0 0
| {z } | {z }
p
2⇡ 1 r2
" 1
#
1 2 3/2 1 2⇡
= 2⇡ (1 r ) = 2⇡ ⇥ = , (4.27)
3 0 3 3
Triple Integrals 49

which is one-half the volume of the unit sphere.

Example. Consider the cone in Fig. 4.7. The surface is given by

x2 + y2 = (1 z)2 , (4.28)

for 0  z  1. The ranges of r and are

0  r  1, 0 < 2⇡ . (4.29)

The range of z is calculated by following the steps in the preceding example. The
cone is bounded from below by the x-y plane, where z = 0. The upper bound of
x is determined by the surface of the cone which, in cylindrical coordinates, is
r2 = (1 z)2 . Thus, the range of z is

0z1 r. (4.30)

The volume integral of the cone is


Z 1 Z 2⇡ Z 1 r Z 1
V = r dr d = 2⇡ (r r2 ) dr
0
|0{z } |{z}
0 0

2⇡ 1 r
!
r2 1 r3 1 1 ⇡
= 2⇡ = 2⇡ ⇥ = . (4.31)
2 0 3 0 6 3

The preceding two examples showed how cyl-


indrical polar coordinates are used to calculate 1
the volumes of surfaces of revolution, i.e. sur-
faces that were obtained by rotating a curve about z
an axes, in those cases, the z-axis. We now con-
1
sider a more substantial example by calculating 1

the volume of another surface of revolution, the 0


0
y
torus. x
11

Example.⇤ A torus is a surface of revolution Fig. 4.7: The surface determined


generated by rotating a circle of radius ⇢ whose by x2 + y2 = (1 z)2 , for 0  z  1.
center is a distance R > ⇢ from the origin about an axis, usually taken as the z-axis.
The calculation of the volume of a torus does not actually require an expression
for the surface. The ranges of r, , and z can be determined by referring to Fig. 4.8.
Consider first the left panel. Suppose we take the range of z as

⇢  z  ⇢. (4.32)
50 Triple Integrals

The equation of the circle in the x-z plane is

(x R)2 + z2 = ⇢2 , (4.33)

so the range of r is obtained by solving this equation for x and referring to Fig. 4.8(b):
p p
R ⇢2 z2  r  R + ⇢2 z2 . (4.34)

z y

x x

(a) (b)
Fig. 4.8: (a) The circle in the x-z plane that is rotated about the z-axis. (b) The section of the
torus in the x-y plane. The emboldened line is the path traced out by the center of the circle.

Figure 4.8(b) indicates that the range of is

0  2⇡ . (4.35)

The volume integral of the torus is thus given by


Z Z Z p 2⇡ ⇢ R+ ⇢2 z2
V= d dz p r dr . (4.36)
0 ⇢ R ⇢2 z2

The radial integral is evaluated as


Z p2 2 R+ ⇢ z 2 R+
p
⇢2 z2
r
p r dr = p
R ⇢2 z2 2 R ⇢2 z2

1 ⇣ p ⌘2 ⇣ p ⌘2
= R + ⇢2 z2 R ⇢2 z2
2
p
= 2R ⇢2 z2 . (4.37)

The integral over the azimuthal angle in Eq. (4.36) is 2⇡, so the volume integral
reduces to Z ⇢p
V = 4⇡R ⇢2 z2 dz . (4.38)

Triple Integrals 51

This integral can be evaluated by the trigonometric substitution z = ⇢ sin ✓, where


1
2
⇡  ✓  12 ⇡. Carrying out the required changes to the integrand, the integration
element, and the limits of integration yields
Z ⇢p Z 12 ⇡
2
V = 4⇡R ⇢ z dz = 4⇡R⇢
2 2 cos2 ✓ d✓ = 2⇡2 R⇢2 . (4.39)
1
⇢ 2⇡
| {z }
1
2

By writing this result as
V = (2⇡R) ⇥ (⇡⇢2 ) , (4.40)
the volume of a torus can be interpreted as the product of the area of the circle that
is rotated about the z-axis to form the torus (⇡⇢2 ) and the length of the path taken
by the center of the circle (2⇡R). This is a special case of Pappus’ Theorem1 : let
R be a planar region that lies entirely on one side of an axis (usually the z-axis) in
the plane. If R is rotated about this axis, the volume of the resulting solid is the
product of the area A of R and the distance travelled by its centroid.

4.3 Spherical Polar Coordinates


One of the most important coordinate systems in physics is spherical polar co-
ordinates. These are appropriate whenever there is a spherical boundary or a sec-
tion of such a boundary. Spherical polar coordinates are especially important for
the quantum mechanical theory of atoms, which is based on spherical symmetry.

4.3.1 Definition of the Coordinate System


In spherical polar coordinates, a point (x, y, z) is expressed in terms of the radius
r, which measures the distance of the point from the origin, the azimuthal angle ,
which measures the orientation of the radius vector with respect to the positive x-
axis, with positive taken in the counterclockwise direction, and the polar angle
✓, which measures the orientation of the radius vector with respect to the z-axis.
These definitions and conventions are depicted in Fig. 4.9.
The transformations between the coordinates (x, y, z) and (r, , ✓) is determined
from the trigonometric construction in Fig. 4.10. The projection of the radial
vector onto the x-y plane has length r sin ✓. The x and y coordinates are obtained
by projecting this quantity onto the x and y axes:
x = r sin ✓ cos , (4.41)
y = r sin ✓ sin . (4.42)
1
Pappus of Alexandria, who lived in the 4th century, is considered to be the last of the great
Greek geometers.
52 Triple Integrals

(a) (b)
Fig. 4.9: Two depictions of spherical polar coordinates. (a) The definitions and ranges of r, ,
and ✓. (b) The representation of any point as the intersection of the surface of constant r (the
sphere), constant (the plane), and constant ✓ (the cone).

The projection of the radius onto the z-axis is

z = r cos ✓ . (4.43)

These are the transformations that relate Cartesian coordinates to spherical polar
coordinates.
The ranges of the radial and azimuthal z
variables are determined by referring to
Fig. 4.9(a). As in circular polar coordin-
ates (Sec. 3.2) r

0  r < 1, 0 < 2⇡ . (4.44)

The range of ✓ is determined by requir- r cos

ing that the transformation between Cartesian


and spherical polar coordinates is single-
r sin
valued, i.e. that one and only one set of
spherical polar coordinates (r, , ✓) cor- Fig. 4.10: The orientation of a radial vector
responds to a particular set of Cartesian with respect to the z-axis.
coordinates (x, y, z). This necessitates restricting the range of ✓ to

0  ✓  ⇡. (4.45)

To understand this, consider a point r

r = r cos sin ✓ i + r sin sin ✓ j + r cos ✓ k . (4.46)

Suppose that we transform this point by rotating the azimuthal angle by ⇡: ! +


⇡. The coordinates of the transformed point r0 are obtained by applying standard
Triple Integrals 53

trigonometric identities:

r0 = r cos sin ✓ i r sin sin ✓ j + r cos ✓ k . (4.47)

Now suppose that we rotate the polar angle of r so that: ✓ ! 2⇡ ✓ (< 2⇡). The
coordinates of the transformed point r00 are again determined by applying standard
trigonometric identities:

r00 = r cos sin ✓ i r sin sin ✓ j + r cos ✓ k . (4.48)

By comparing these coordinates with those in Eq. (4.47), we conclude that r00 = r0 ,
i.e. that there are two ways of labelling the same point. To avoid this unacceptable
result, the range of ✓ is restricted to the range in Eq. (4.45).

4.3.2 The Integration Element


The integration element in spherical polar coordinates is most easily obtained with
the procedure in Problem Set 4. The radius vector associated with a point (x, y, z)
is written as
r = r cos sin ✓ i + r sin sin ✓ j + r cos ✓ k , (4.49)

and calculating the integration element from the vector product

dV = |drr ·dr ⇥ dr✓ | , (4.50)

where drr , dr , and dr✓ are the di↵erential changes of r with respect to r, , and
z, respectively:

drr = dr cos sin ✓ i + dr sin sin ✓ j + dr cos ✓ k , (4.51)

dr = r sin sin ✓ d i + r cos sin ✓ d j , (4.52)

dr✓ = r cos cos ✓ d✓ i + r sin cos ✓ d✓ j r sin ✓ d✓ k . (4.53)

These vectors are mutually orthogonal so the integration element is obtained from
the product of their magnitudes:

dV = r2 sin ✓ dr d d✓ . (4.54)
54 Triple Integrals

4.3.3 Triple Integrals in Spherical Polar Coordinates


Integrals of a function F(x, y, z) over a volume V are written in spherical polar
coordinates as
ZZZ
F(x, y, z) dx dy dz
V
ZZZ
⇥ ⇤
= F x(r, , ✓), y(r, , ✓), z(r, , ✓) r2 sin ✓ dr d d✓
V0
ZZZ
⌘ f (r, , ✓) r2 sin ✓ dr d d✓ , (4.55)
V0

where V 0 is the volume V expressed in spherical polar coordinates. There are two
important special cases of this integral. If f has no -dependence, f = f (r, ✓),
then f is said to have azimuthal symmetry. According to the transformations in
Eqs. (4.42) and (4.43) and Fig. 4.8, this corresponds to rotational symmetry about
the z-axis. Surfaces of revolution have this type of symmetry. A physical situation
with this type of symmetry is discussed in Problem Set 4. The integral over can
be evaluated immediately and the general expression in Eq. (4.55) becomes
ZZ
2⇡ f (r, ✓) r2 sin ✓ dr d✓ . (4.56)

In the second case, where f has neither - nor ✓-dependence, f is said to be


isotropic. This corresponds to “spherical” symmetry in that f depends only on
the radius r and not on any angular orientation. The integrals over and ✓ can be
evaluated immediately and the general integral in Eq. (4.55) reduces to
Z
4⇡ f (r)r2 dr . (4.57)

This integral is seen to correspond to the integration over radial shells.

Example. We consider first the calculation of the volume of a sphere of radius


R. Referring to Eq. (4.55), this corresponds to the case f = 1. The ranges of the
integration variables are obtained directly from Fig. 4.9(a):

0  r  R, 0 < 2⇡ , 0  ✓  ⇡, (4.58)

so the volume integral is


Z R Z 2⇡ Z ⇡
2 4
V= r dr d sin ✓ d✓ = ⇡R3 . (4.59)
3
|0 {z } |0{z } |0 {z }
1 3
R 2⇡ 2
3
Triple Integrals 55

The generalization of this procedure to sections of a sphere between given azi-


muthal and polar angles and to spherical shells with given inner and out radii is
straightforward.

Example. Consider the integral of f = e ↵r over all space. This is an ex-


ample of a function with spherical symmetry that occurs frequently in quantum
mechanics. The ranges of the integration variables are

0  r < 1, 0 < 2⇡ , 0  ✓  ⇡, (4.60)

so the integral of f becomes


Z 1 Z 2⇡ Z ⇡ Z 1
2
re ↵r
dr d sin ✓ d✓ = 4⇡ r2 e ↵r
dr . (4.61)
0
|0{z } |0 {z } 0

2⇡ 2
The radial integral is evaluated by performing successive integrations by parts:
Z 1 Z
2 1 2 ↵r 1 8⇡ 1 ↵r
4⇡ r e| {zdr
|{z}
↵r
} = ↵r e 0 + ↵ re dr
0 u | {z } 0
dv
0
Z
8⇡ ✓ 1 ↵r 1 1 1 ↵r ◆
= re + e dr
↵ | ↵ {z } 0 ↵ 0
0
!
8⇡ 1 ↵r 1
= 2 e
↵ ↵ 0

8⇡
= . (4.62)
↵3
Notice that, in arriving at this result, we have twice used the fact that

lim xn e ↵x
=0 (4.63)
x!1

for any n (and ↵ > 0).

4.4 Surface Integrals


A particular case of integrals in three dimensions involves integrals over surfaces.
A common type of surface integral is where one of the three variables is held
56 Triple Integrals

constant. Consider the surface of a sphere of radius R. According to Eq. (4.49)


the radius r vector at any point on the sphere is

r = R cos sin ✓ i + R sin sin ✓ j + R cos ✓ k , (4.64)

The element of area integration is obtained by calculating the di↵erential of this


vector for changes in turn of d and d✓:

dr = R sin sin ✓ d i + R cos sin ✓ d j , (4.65)

dr✓ = R cos cos ✓ d✓ i + R sin cos ✓ d✓ j R sin ✓ d✓k . (4.66)

These vectors are orthogonal,

dr ·dr✓ = 0 , (4.67)

so the di↵erential area dA corresponding to these di↵erential changes is obtained


from the product of the magnitudes of dr and dr :

dA = |dr ||dr✓ | = R sin ✓ d ⇥ R d✓ = R2 sin ✓ d d✓ . (4.68)

Example. The surface area of a sphere of radius R is represented as


Z 2⇡ Z ⇡
2
R d sin ✓ d✓ = R2 ⇥ 2⇡ ⇥ 2 = 4⇡R2 . (4.69)
0 0

The corresponding expression of the surface area subtended by azimuthal angles


1 and 2 and polar angles ⇥1 and ⇥2 is
Z 2 Z ⇥2
2
R d sin ✓ d✓ = R2 ( 2 1 )(cos ⇥1 cos ⇥2 ) . (4.70)
1 ⇥1

The other type of surface integral we will encounter involve a cylinder of ra-
dius R. The radius vector is, from Eq. (4.19), given by

r = R cos i + R sin j + z k, (4.71)

The di↵erential dr corresponding to di↵erential changes of d and dz are

dr = R sin d i + R cos d j , (4.72)

drz = dz k . (4.73)
Triple Integrals 57

These vectors are manifestly orthogonal, dr ·drz = 0, so the di↵erential area dA


corresponding to these di↵erential changes is obtained from the product of the
magnitudes of dr and drz :
dA = R d dz . (4.74)

Example. The surface of a cylinder of radius R and height H is calculated as


Z 2⇡ Z H
R d dz = 2⇡RH . (4.75)
0 0

The surface area of cylinder between heights H1 and H2 and azimuthal angles 1
and 2 is similarly calculated as
Z 2 Z H2
R d dz = R( 2 1 )(H2 H1 ) . (4.76)
1 H1

4.5 Summary
The triple integral of a function f (x, y, z), viewed as a density of some physical
quantity, is the amount of that quantity within a volume in three-dimensional
space. There is considerably more freedom to specify other cooordinate systems
than in two dimensions and many applications in physics rely on such transform-
ations to enable calculations to be carried out. From Cartesian coordinates, we
transformed triple integrals into cylindrical polar coordinates, which are the nat-
ural generalizations of circular polar coordinates to three dimensions, and are ap-
propriate to situations where there is azimuthal symmetry, and spherical polar co-
ordinates, for situations that involve spherical symmetry. The Jacobians obtained
in each case reflect the position dependence of the magnitude of the di↵erential
volume elements.
Chapter 5

Line Integrals

For a function f of a single variable, the integral of f over an interval is uniquely


determined once the limits of integration are specified. Extending this construc-
tion to the integration of a function f of two or more variables along a path in
space connecting specified initial and final points (the limits of integration) leads
to entirely new mathematical issues. Foremost among these is that the value of
such an integral – called a line integral – generally depends not just on the limits of
integration, but on the path that connects these points along which the integration
of f is carried out. Thus, the information required to perform a line integral of a
given function is comprised of the initial and final points and the path connecting
them. In this respect line integrals represent a significant conceptual departure
from double and triple integrals.
In this chapter, we will first motivate the mathematical structure of a large
class of line integrals, using the calculation of work in classical mechanics as
a motivation, and work through several examples to demonstrate through explicit
calculations that the value of a line integral can depend on the path connecting two
points. We will then examine some general properties of line integrals, determine
the criterion for the value of a line integral to be independent of the integration
path between the limits of integration. Path-dependent and path-independent line
integrals each have important applications in several areas of physics, including
mechanics, electromagnetic theory, and thermodynamics, which we mention at
various places in this chapter.

5.1 Work in Classical Mechanics


A standard calculation in classical mechanics is the work W done by a force
F along a path between two points a and b. If the force has a constant magnitude
and direction and acts at an angle ✓ along a path of length r, as shown in the figure
at right, the work W done by the force is W = F· r. Similarly, if F acts only over

59
60 Line Integrals

an infinitesimal distance dr, the corresponding work dW done is


dW = F·dr . (5.1)
Suppose now that the force is a func- F
tion of position. We consider this situ-
ation in one dimension first: F = F(x).
The calculation of the work between two
points x = a and x = b proceeds accord-
ing to the construction in Fig. 5.2. The in- r
terval (a, b) is first divided into N subin-
tervals of length x = (b a)/N. The Fig. 5.1: A force F acting at an angle ✓
force acting within each of these subin- along a displacement r.
tervals is taken to be the constant value at the left endpoint of that interval. Thus,
we obtain
F(a) x + F(a + x) x + · · · + F(b x) x . (5.2)
as an approximation of the work done over the interval. As x ! 0, this approx-
imation becomes increasingly accurate and the work done approaches the shaded
region in the right panel of the figure. Referring to Sec. 2, the procedure depic-
ted in Fig. 5.2 is the same as that used for the Riemann sum construction of the
integral of a function, so we conclude that
Z b
W= F dx . (5.3)
a

Similar considerations apply for paths in two and three dimensions. In this
chapter, we will consider the two-dimensional case. A force F in two dimensions

F F

x x
a b a b

Fig. 5.2: (Left panel) Construction used to calculate the work done from x = a to x = b by a
position-dependent force. The shaded area corresponds to the work calculated by regarding
the force as constant over each subinterval. (Right panel) The corresponding calculation for
infinitesimal subintervals, which is seen to represent the area bounded by F , the x-axis, and
the lines x = a and x = b.
Line Integrals 61

is a vector field:
F(x, y) = P(x, y) i + Q(x, y) j , (5.4)
where P and Q are functions of x and
y and i. This expression indicates that f
every point (x, y) is assigned a vector F
whose x-component is given by P i and
whose y-component is Q j. The path along
which F acts is a curve P in the x-y plane
between an initial point i and a final point
f , as shown in Fig. 5.3. The work done
along this path is calculated as in Eq. (5.1)
i
by first considering the incremental work
dW done by the force along a distance
dr: dW = F · dr, where dr is the incre-
mental distance along the path. Then, Fig. 5.3: A path in a vector field between an
with the position vector given by initial point i and the final point f .

r = xi + y j, (5.5)

we have that the incremental change along the path is

dr = dx i + dy j , (5.6)

so the work done along P is


Z
W = F·dr
P
Z
⇥ ⇤
= P(x, y) i + Q(x, y) j ·(dx i + dy j)
P
Z
⇥ ⇤
= P(x, y) dx + Q(x, y) dy . (5.7)
P

This is an example of a line integral.


In addition to this example from classical mechanics, line integrals appear in
thermodynamics and in electricity and magnetism. In thermodynamics, P rep-
resents a process between initial and final values of thermodynamic variables
(e.g. pressure, temperature, volume). The line integral of such variables yields
quantities such as heat flow and the work done during the process. In electricity
and magnetism, P is a path in space, and line integrals represent quantities such
as the electromotive force. In all of these cases the mathematical form of a line
62 Line Integrals

integral is
Z
⇥ ⇤
f (x, y) dx + g(x, y) dy , (5.8)
P

where f and g are any functions that can be integrated and P is the path connecting
the initial and final points.
As we stressed in the introduction, specifying the integration path P is as
important as specifying the initial and final points. The path provides a functional
relationship between x and y and allows the integrals to be evaluated; otherwise
the variable y in the term f (x, y) dx and the variable x in the term g(x, y) dy appear
superfluous. Additionally, the value of the line integral may depend explicitly
on the path, so specifying only the initial and final points does not necessarily
sufficient to obtain a unique value. The following example illustrates these ideas.
Example. Consider the line integral
Z
xy dx , (5.9)
P
which is of the general form in Eq. (5.8) with f = xy and g = 0. In the context of
the calculation of work, this corresponds to a force F(x, y) = xy i.
We will evaluate this integral over 1
the three paths shown in Fig. 5.4, each
of which have their initial point at the 0.8
origin (0, 0) and their final point at
(1, 1). 0.6 P2
We first consider P1 . This path is
y

composed of two straight segments: 0.4


P3
(0, 0) ! (1, 0) and (1, 0) ! (1, 1).
0.2
The first segment lies along the x-axis,
so we have that P1
0
y = 0, 0  x  1. (5.10) 0 0.2 0.4 0.6 0.8 1
x
Hence, since y = 0, the integrand van-
Fig. 5.4: The three paths, labelled P1 , P2 , and
ishes, so the contribution from seg-
P3 between (0, 0) and (1, 1) used for evaluating
ment also vanishes. The second seg-
the line integral in Eq.(5.9).
ment is parallel to the y-axis, so
x = 1, dx = 0 , 0  y  1. (5.11)
Since dx = 0, the contribution along this segment also vanishes. Therefore, the
integral along P1 vanishes: Z
xy dx = 0 . (5.12)
P1
Line Integrals 63

The path P2 connects (0, 0) to (1, 1) with the straight line y = x. Thus, along
this path, the integrand can be expressed entirely as a function of x: xy = x2 , with
0  x  1. The line integral is thereby evaluated as
Z Z 1 1
xy dx = x2 dx = 31 x3 = 13 . (5.13)
P2 0 0

Finally, the path P3 connects (0, 0) to (1, 1) with the parabola y = x2 . Along
this path, the integrand can be written as xy = x3 , with 0  x  1, and the line
integral becomes
Z Z 1 1
xy dx = x3 dx = 14 x4 = 14 . (5.14)
P3 0 0

We have thus obtained three di↵erent values for the line integral in Eq. (5.9)
along the three paths shown in Fig. 5.4. This result can be understood by inter-
preting this integral as the work done by the force F = xy i over the three paths:
Z Z
F·dr = xy dx , (5.15)
Pi Pi

for i = 1, 2, 3. This vector field is shown 0 0.2 0.4 0.6 0.8 1


in Fig. 5.5 superimposed on the paths P1 , 1
P2 , and P3 . We can see immediately from
this diagram that the work done along P1 0.8
must vanish because F vanishes along the
0.6 P2
x-axis (the first segment of P1 ), and acts in
the normal direction to the second segment
y

0.4
of this path. Alternatively, the line integrals P3
along P2 and P3 are both necessarily pos- 0.2
itive because the projection of F onto the
P1
path has a component along the direction 0
of the path, producing positive work. 0 0.2 0.4 0.6 0.8 1
x

This example illustrates two fundamental Fig. 5.5: The vector field F = xy i and the
points about line integrals. (i) The value of paths P1 , P2 , and P3 shown in Fig. 5.4
a line integral may depend on the path over used for the evaluation of the line integral
which it is evaluated. There are physical in Eq. (5.9).
manifestations of this property that have im-
portant consequences in mechanics, thermodynamics, and electricity and magnet-
ism. (ii) The path between given initial and final points establishes a relationship
between the independent variables. Once this information is incorporated into the
line integral, the evaluation reduces to that of an ordinary integral (Sec. 2).
64 Line Integrals

Example. Consider the line integral


Z
(xy2 dx + x2 y dy) (5.16)
P

evaluated along the three paths in Fig. 5.4. Along the first segment of P1 , y = 0,
and therefore dy = 0, so there is no contribution from either term in the integral.
Along the second segment x = 1, dx = 0, and 0  y  1. Thus, only the second
term in the integral makes a contribution to the integral, and we obtain
Z Z 1 1
2 2
(xy dx + x y dy) = y dy = 12 y2 = 12 . (5.17)
P1 0 0

Along P2 , y = x, so dy = dx. Thus, both terms in the integrand can be written


in terms of either x or y alone:
R R1
P2
(xy2 dx + x2 y dy) = 2 0
x3 dx
1
= 2 ⇥ 14 x4 = 12 , (5.18)
0

which is the same value obtained in Eq. (5.17).


Finally, along P3 , y = x2 , so dy = 2x dx. We can express the integrand in
terms of x alone to obtain
R R1
2 2
P3
(xy dx + x y dy) = 0
(x5 dx + 2x5 dx)
R1 1
= 3 0 x5 dx = 3 ⇥ 16 x6 = 12 , (5.19)
0

which is the same as that obtained for the other two paths. A natural question
arises: Is this a coincidence, or does this integral always have the same value when
evaluated over di↵erent paths between fixed initial and final points? The results
we have obtained in this example are certainly suggestive, but to address this
question in a mathematically concise framework, we must derive some additional
properties of line integrals. This is the subject of the next two sections.

5.2 Line Integrals over Closed Curves


An important class of line integrals is that for which the path of integration forms
a simple closed curve, i.e. a path that returns to the initial point but does not
cross its path. Such integrals find many applications in thermodynamics, where
they are called “cycles”, and in electricity and magnetism, where they form the
mathematical expression of Ampère’s law, and are referred to as “loop integrals”.
Line Integrals 65

In this section, we will re-express the question of the path-dependence of a line


integral in terms of the value of that integral around a closed curve. We first
determine the e↵ect that reversing the sense of the integration path has on the
value of a line integral.
Consider a line integral over a path between an initial point i and a final point
f , as shown in Fig. 5.6(a):
Z y

( f dx + g dy) . (5.20) f
P
Suppose that this path is reversed, so that the new
P
initial point is f and the new final point is i, as shown
Fig. 5.6(b). We signify this path by P and write the
corresponding line integral as i
Z (a)
( f dx + g dy) . (5.21) x
P
The relationship between the values of these two line y

integrals is straightforward to understand. As the ex- f


amples in the preceding section show, the evaluation of
a line integral always reduces to an ordinary integral. P
Thus, reversing the integration path in a line integral
has the e↵ect of interchanging the upper and lower lim-
its of integration. According to the Fundamental The- i
(b)
orem of Calculus, this changes the sign of the integral
x
[Eq. (2.15)]. Thus, the line integrals in Eqs. (5.20) and
(5.21) have the same absolute value, but opposite signs: Fig. 5.6: (a) The path P
between points i and f ,
Z Z and (b) the reverse path
( f dx + g dy) = ( f dx + g dy) . (5.22) P (b).
P P

Consider now a line integral over a closed curve C (Fig. 5.7). Such integrals,
often called “loop integrals”, have a special notation to indicate that the integration
path is a closed curve: I
( f dx + g dy) . (5.23)
C
Choose any two distinct points A and B on C and denote by P1 the path on C from
A to B and by P2 the path that returns B to A along C. The integral over C can be
expressed as sum of line integrals over P1 and P2 :
I Z Z
( f dx + g dy) = ( f dx + g dy) + ( f dx + g dy) . (5.24)
C P1 P2
66 Line Integrals

Suppose that the value of the line integral in Eq. (5.20) is independent of the path
P for any initial and final points. The closed curve C in Fig. 5.7 defines two
paths from A to B: the path P1 and the reverse of the path P2 . Path-independence
requires that the line integrals over P1 and P2 are equal:

Z Z
( f dx + g dy) = ( f dx + g dy) . (5.25)
P1 P2

By invoking Eq. (5.22), we can write this equation as

Z Z
( f dx + g dy) ( f dx + g dy)
P1 P2
Z Z
= ( f dx + g dy) + ( f dx + g dy)
P1 P2
I
= ( f dx + g dy) = 0 . (5.26)
C

This shows that, if the value of a line integral is independent of the path between
any initial and final points, the loop integral vanishes for any closed curve.
The converse of this statement is also y
true. If a loop integral vanishes for any
closed curve C, then we can choose any P1
two points A and B on C as initial and fi-
nal points of line integrals along the cor-
responding paths P1 and P2 . Then, by re-
versing the steps leading to Eq. (5.26), we B

find that
Z Z
A
( f dx + g dy) = ( f dx + g dy) , P2
P1 P2
(5.27) x

which implies path independence. Thus, Fig. 5.7: A closed curve C in the x-y plane.
we have shown that the path independ- P1 is a path between any two points A and
ence of a line integral is both necessary B on C and P2 is the path from B to A that
[Eq. (5.27)] and sufficient [Eq. (5.26)] for completes the loop. The closed curve is the
the loop integral to vanish over any closed sum of these two paths: C = P1 + P2 .
curve. In other words, these two proper-
ties are equivalent:
Line Integrals 67

A line integral
Z
( f dx + g dy)
P

is independent of the path P between any two points i and f if and only if
I
( f dx + g dy) = 0
C

for any closed curve C.

This result provides an alternative statement of the fact that line integrals fall into
two classes: (i) path-dependent and, therefore, typically non-vanishing values over
closed curves, and (ii) path-independent and vanishing values over closed curves.
Both types of line integral are important in applications to physics and under-
standing the physical circumstances that lead to one type of integral or another
is a central theme in several disciplines. We conclude this section with two ex-
amples.

Example. Consider the loop integral


I y

y dx , (5.28)
C
where C is a circle of radius a centered at
(1, 1), as shown in Fig. 5.8. We represent
a
the circle as follows: 1

x=1 a cos ,
(5.29)
y = 1 + a sin ,
where 0  < 2⇡. This parametrization
x
sweeps through the circle in a clockwise 1
direction beginning at (1 a, 1). The in-
Fig. 5.8: The circle of radius a centered at
tegral in Eq. (5.28) can be expressed as
(1, 1), showing the definition of for carry-
an integral over by using Eq. (5.29) to
ing out a loop integral over this curve.
transform the integrand, the integration ele-
ment, and the limits of integration. The integrand y is given by the second of
Eqs. (5.29), an application of the chain rule to x( ) yields
dx = a sin d , (5.30)
68 Line Integrals

y y
(a) (b)
P1

B
A

P2
xA xB
x x

y y
(c) (d)

x x

Fig. 5.9: The evaluation of the loop integral in Eq. (5.28) around an arbitrary closed curve
C, showing (a) the separation of C into upper and lower paths P1 and P2 , (b) and (c) the
evaluation of the integral along these paths, and (d) the cumulative effect of the loop integral.

and the limits of integration are 0  < 2⇡. The original integral thereby becomes
I Z 2⇡
y dx = (1 + a sin )a sin d
0
Z 2⇡ Z 2⇡
=a sin d +a 2
sin2 d
|0 {z } |0 {z }
=0 =⇡
= ⇡a2 . (5.31)

This is readily identified as the area of the circle enclosed by C.

This result can be generalized to any closed curve in the x-y plane by following
the steps shown in Fig. 5.9. We first identify the points A = (xA , yA ) and B =
(xB , yB ) that allow C to be written as the sum of upper and lower paths P2 and P2 ,
which can be represented as functions y1 (x) and y2 (x), respectively. The integral
Line Integrals 69

along P1 is Z Z xB
y dx = y1 (x) dx . (5.32)
P1 xA

This is an ordinary integral whose value is represented by area bounded by y1 (x),


the x-axis, and x = xA and x = xB , as shown in Fig. 5.9(a). The loop is completed
by integrating y2 (x) from xB to xA . The integral
Z xB
y2 (x) dx (5.33)
xA

is represented by the area shown in Fig. 5.9(c). But the integral we need to com-
plete C has the upper and lower limits interchanged, so its value corresponds to
the negative of this quantity. Hence, the loop integral is calculated as
I Z xB Z xB
y dx = y1 (x) dx y2 (x) dx , (5.34)
C xA xA

which is represented in Fig. 5.9(d). The integral over y2 (x) cancels the contribution
from the integral over y1 (x) that represents the area below P2 , leaving only the area
enclosed by C. We have thereby shown that
I
y dx = A , (5.35)
C

where A is the area enclosed by C.

We conclude this section with an ex- y


ample of a loop integral that does vanish.

( 1,1) (1,1)
Example. Consider the integral
I
(xy2 dx + x2 y dy) , (5.36) x
C

where C is the closed curve in Fig. 5.10.


The integrand is the same as that in the ( 1, 1) (1, 1)
second example in Sec. 5.1. The closed
curve is composed of four straight seg-
ments, so we will evaluate the loop in- Fig. 5.10: The closed contour for the integ-
tegral by considering each segment sep- ral in Eq. (5.36).
arately. Beginning at ( 1, 1), the seg-
ments are characterized as follows:
70 Line Integrals

( 1, 1) ! ( 1, 1) : x= 1 dx = 0 1y1
( 1, 1) ! (1, 1) : y=1 dy = 0 1x1
(1, 1) ! (1, 1) : x=1 dx = 0 1y1
(1, 1) ! ( 1, 1) : y= 1 dy = 0 1x1

If dx = 0 the first term in Eq. (5.36) makes no contribution, while if dy = 0, the


second term makes no contribution. The integral can therefore be written as (note
the upper and lower limits of each integral!)
Z 1 Z 1 Z 1 Z 1
y dy + x dx + y dy + x dx = 0 . (5.37)
1 1 1 1

5.3 Exact and Inexact Di↵erentials


Although the results of the preceding section allow us to re-express the path-
independence of a line integral in terms of a loop integral, we are no closer to
determining a priori whether or not a given line integral is independent of path
between given initial and final points. In this section, we will derive a condition
that allows us to address this question without having to perform any integration
whatsoever.
Consider the line integral
Z
( f dx + g dy) (5.38)
P

over a path P between an initial point (xi , yi ) and a final point (x f , y f ). The path
establishes a relation between x and y that we represent as y(x). This enables us
to write the line integral as an integral over x only by following the procedure in
Sec. 2.2. We have
Z Z xf
f (x, y) dx = f [x, y(x)] dx , (5.39)
P xi
Z Z xf
dy
g(x, y) dy = g[x, y(x)] dx . (5.40)
P xi dx
Thus, Z Z ( )
xf
dy
( f dx + g dy) = f [x, y(x)] + g[x, y(x)] dx . (5.41)
P xi dx
The right-hand side of this equation is an ordinary integral over xi  x  x f .
Accordingly, if the line integral is path-independent, we can use the Fundamental
Line Integrals 71

Theorem of Calculus to write


Z xf ( )
dy
f [x, y(x)] + g[x, y(x)] dx = F(x f ) F(xi ) , (5.42)
xi dx

where
dF dy
= f [x, y(x)] + g[x, y(x)] . (5.43)
dx dx
By writing F as F[x, y(x)], we also have

dF @F @F dy
= + , (5.44)
dx @x @y dx
from which we identify
@F @F
= f, = g. (5.45)
@x @y
The quantity F is called the potential. On account of Eq. (5.45) we can write the
di↵erential of F as
@F @F
dF = dx + dy = f dx + g dy , (5.46)
@x @y
in which case the quantity on the right-hand side is independent of the path. This
is called an exact differential. Otherwise, the quantity f dx + g dy is called an
inexact differential and the corresponding line integral is path-dependent. Hence,
a line integral of an exact di↵erential can be represented as
Z Z
( f dx + g dy) = dF , (5.47)
P P

in which case we have that


Z
( f dx + g dy) = F[x f , y(x f )] F[xi , y(xi )]
P

= F(x f , y f ) F(xi , yi ) , (5.48)

In terms of our original formulation in Sec. 5.1, this equation states the work done
between and initial point i and a final point f is equal to the change in the potential
F.
Equation (5.45) provides a method of testing for the exactness of a di↵erential.
By di↵erentiating the first of these equations with respect to y,
!
@ @F @2 F @f
= = , (5.49)
@y @x @y@x @y
72 Line Integrals

the second with respect to x,


!
@ @F @2 F @g
= = , (5.50)
@x @y @x@y @x
and equating the mixed second partial derivatives of F: Fyx = F xy , we obtain

@f @g
= . (5.51)
@y @x

The discussion leading to this equation shows that it is a necessary condition for
a di↵erential to be exact. The procedure described in Problem 5, Problem Set 5
shows that this is also a sufficient condition for exactness, thus demonstrating the
equivalence between Eq. (5.51) and the exactness of a di↵erential.

Example. Consider the line integral


Z
y dx , (5.52)
P

which was discussed in an earlier Example in this section. In the notation of


Eq. (5.8), f = y and g = 0. Thus,
@f @g
= 1, = 0. (5.53)
@y @x
Since these two partial derivatives are unequal, we conclude from Eq. (5.51) that
y dx is an inexact di↵erential, so the line integral in Eq. (5.52) is path-dependent.
This result is to be expected in view of Eq. (5.35).
As a second example, we consider the line integral
Z
(xy2 dx + x2 y dy) . (5.54)
P

This integral has been discussed in Secs. 5.1 and 5.2. In the notation of Eq. (5.8),
f = xy2 and g = x2 y, and we find
@f @g
= 2xy , = 2xy . (5.55)
@y @x

The equality of these partial derivatives means that xy2 dx + x2 y dy is an exact


di↵erential, so the line integral in Eq. (5.54) is path-independent. This conclu-
sion confirms our expectations based on the results already obtained for this line
integral.
Line Integrals 73

The exactness of this di↵erential implies that there is an underlying potential


function F such that
@F @F
= xy2 , = x2 y . (5.56)
@x @y
Integrating the first of these with respect to x yields

F(x, y) = 12 x2 y2 + h(y) , (5.57)

where h(y) is an arbitrary function of y (analogous to constants of integration


obtained when integrating functions of one variable). Di↵erentiating both sides of
this equation with respect to y,
@F
= x2 y + h0 (y) , (5.58)
@y
and requiring that this result be consistent with the second of equations (5.56),
necessitates setting h = constant, so h0 (y) = 0. Thus,

F(x, y) = 12 x2 y2 + constant . (5.59)

The constant term disappears upon integration:


Z
(xy2 dx + x2 y dy) = 12 (x2f y2f xi2 y2i ) . (5.60)
P

5.4 Arc Length⇤


A line integral with a mathematical structure di↵erent from that in Eq. (5.8) is ob-
tained by calculating the distance travelled by a particle moving along a trajectory.
The trajectory is described by a curve [x(t), y(t)], where t is the time between an
initial time ti and a final time t f . The distance ds travelled by the particle in time
dt, when x changes by dx and y changes by dy, is given by the relation

ds2 = dx2 + dy2 , (5.61)

or,
ds = (dx2 + dy2 )1/2 . (5.62)
Thus, the total distance S travelled by the particle, called the arc length of the
path, is Z f Z f p
S = ds = dx2 + dy2 . (5.63)
i i
74 Line Integrals

The integrand on the right-hand side can be written in a more physically suggest-
ive form as
2 ! !2 31/2 q
p 666 dx 2 dy 777
dx2 + dy2 = dt 46 + 75 = dt v2x + v2y = v dt , (5.64)
dt dt

where v x and vy are x and y components of the instantaneous speed v of the particle.
The arc length along the trajectory is can thereby be represented as
Z tf
S = v(t) dt . (5.65)
ti

The general form of the arc length,

Z p
dx2 + dy2 . (5.66)
P

is used to represent the distance along any curve P.

Example. We will illustrate the methodology of computing the arc length by


considering y = cosh x between x = 0 and x = a. For y = cosh x, we have

dy = sinh x dx , (5.67)

so the integrand in Eq. (5.66) becomes


p p p
dx2 + dy2 = dx2 + sinh2 x dx2 = dx 1 + sinh2 = cosh x dx . (5.68)

Thus, Z p Z a a
dx2 + dy2 = cosh x dx = sinh x = sinh a . (5.69)
0 0

5.5 Summary
We can summarize the main results we have obtained on line integrals by noting
that the following statements are equivalent in that any one implies any other. If
any one statement is false, all other are false as well.

1. f dx + g dy is an exact di↵erential;
Line Integrals 75
Z
2. ( f dx + g dy) is independent of the path P between fixed endpoints;
P
I
3. ( f dx + g dy) = 0 for any closed curve C;
C

@f @g
4. = ;
@y @x
5. There is a potential function F such that F x = f and Fy = g, so dF =
f dx + g dy;
Z
6. ( f dx + g dy) = F(x f , y f ) F(xi , yi ) for any initial point (xi , yi ) and final
P
point (x f , y f ).
1st year E & M Key formulae 1 Peter Török

1 Field and potential


1.1 Coulomb’s law (1785)
Force on charge q due to charge Q
Qq
F= r̂ [N] (1.1)
4πε0 r 2
where r̂ points from Q to q, and r is the separation of the charges.

1.2 Electric field of a point charge Q


F
E= or F = qE (1.2)
q
Q
E= r̂ [N/C] = [V/m] (1.3)
4πε0 r 2

1.3 Electrostatic potential energy of a point charge


Qq
U= [J] (1.4)
4πε0 r

1.4 Electrostatic potential of a point charge


U
V= [V] = [J/C] (1.5)
q
Q
V= (1.6)
4πε0 r

1.5 Links between F and U , and between E and V


dU dU
Fr = − or F = − r̂ [N] = [J/m] (1.7)
dr dr
dV dV
Er = − or E = − r̂ [V/m] (1.8)
dr dr
dU = −F · d l dV = −E · d l (1.9)
where Fr and Er are respectively the Coulomb force and the electric field in the direction of increas-
ing r .
The force points in the direction of DECREASING potential energy. The electric field points in the
direction of DECREASING electrostatic potential.

1
1.6 Potential difference
!
∆V ≡ VB − VA = − E.d l (1.10)
A →B
"
E.d l = 0 (1.11)
A →B

The integral in Eq. (1.10) is known as a line (or path) integral. When the path is closed, as in
Eq. (1.11), it is known as a loop integral.

1.7 Relationship between E and V in Cartesian co-ordinates and vector notation


∂V ∂V ∂V
Ex = − ; Ey = − ; Ez = − . (1.12)
∂x ∂y ∂z
# $
∂V ∂V ∂V
E=− i +j +k = −∇V (1.13)
∂x ∂y ∂z

1.8 Conservative fields


The following are equivalent signatures of a conservative field:
%
• E · d l is independent of path

• the loop integral is zero (Eq. (1.11));

• E = −∇V (Eq. (1.13), or the component relations Eq. (1.12)).

1.9 Principle of superposition


Force on q due to Q1 and Q2 :
 
q  Q1 r̂1 Q2 r̂2 
F=  + 2  = F1 + F2 (1.14)
4πε0 r12 r2

Electric field of Q1 and Q2 at given location


 
 Q1 r̂1 Q2 r̂2 
1
E=  + 2  = E1 + E2 (1.15)
4πε0 r12 r2

Electric potential of Q1 and Q2 at given location


# $
1 Q1 Q2
V= + = V1 + V2 (1.16)
4πε0 r1 r2

1.10 Electric potential energy of 3 charges


# $
1 Q1 Q2 Q2 Q3 Q1 Q3
U= + + (1.17)
4πε0 r12 r23 r13

2
1.11 Potential and field on the axis of a dipole
p
V= ; (1.18)
4πε0 x 2
∂V p
Ex ≡ − = (1.19)
∂x 2πε0 x 3
where dipole moment p = Qa with a being the charge separation. When defined as a vector, p
points from − to +.

1.12 Off axis dipole field


The x -component is
p
Ex = (3 cos2 θ − 1) (1.20)
4πε0 r 3
See Question 6 on Problem Sheet 1 for the y - and z -components.

1.13 Potential energy U of a dipole in a uniform electric field


U = −p · E (1.21)

1.14 Torque Γ on a dipole in a uniform electric field


Γ=p×E (1.22)

1.15 Dipole moment of a general charge distribution


,
pn = Qi ni , n = x, y, z (1.23)
i
,
p = px i + py j + pz k = Qi ri (1.24)
i
-
If Qi = 0 (i.e. zero net charge), p is independent of the choice of origin.
i

2 Lines, flux and Gauss’s flux law


2.1 Element of electric flux
d ΦE = E · d S (2.1)

2.2 Total flux through a surface


! !
ΦE = E · dS = E⊥ dS (2.2)
S S

These objects are known as surface integrals. Note that a single (rather than a double) integral sign
is sometimes used.

2.3 Gauss’s flux law


"
1 ,
E · dS = Qi (2.3)
ε0 i
S

The flux through a closed surface = the total charge enclosed ×ε−0 1 .

3
3 Electric field calculations
3.1 Foundation principle
The following formulae for the field and potential always work, although in simple cases one can
use Gauss’s Flux Theorem to avoid using them.

1 , Qi r̂i
E= (3.1)
4πε0
i
ri2

1 ,Q
i
V= (3.2)
4πε0 ri
i

3.2 Spherical shell


radius a , total charge Q
Q



 4πε0 r 2
(r > a )

E= (3.3)



0 (r < a )

3.3 Spherical distribution


radius a , uniform charge density ρ

ρa 3 Q




 3ε0 r 2
= 4πε0 r 2
(r > a )
E= (3.4)


 ρr
(r < a )


3ε0

3.4 Infinite line


linear charge density λ, at distance r
E = λ/2πε0 r (3.5)

3.5 Infinite plane sheet


areal charge density σ
σ
E= (3.6)
2ε0

3.6 Circular ring


linear charge density λ, radius a , at distance z on axis

aλ z aλ Q
E= → 2
≡ as z → ∞ (as it should)! (3.7)
2 2
2ε0 (a + z ) 3/2 2ε0 z 4πε0 z 2

3.7 Circular disk


areal charge density σ, radius a , at distance z on axis
# $
σ z σ
E= 1− √ → as z → 0 (as it should)! (3.8)
2ε0 z2 + a2 2ε0

4
3.8 Outside an infinite plane conductor
areal charge density σ
σ
E= (3.9)
ε0

3.9 Inside a conductor under static conditions


E = 0; V = constant (3.10)

4 Capacitance
C = Q /V (4.1)

4.1 Isolated sphere


radius a
C = 4πε0 a (4.2)

4.2 Two concentric spheres


radii a and (b > a ) # $
b
C = 4πε0 a (4.3)
b −a

4.3 Parallel plate capacitor


plate area A , plate separation s
ε0 A
C= (4.4)
s

4.4 Stored energy in capacitor


1 Q2 1
U= = CV 2 (4.5)
2 C 2

4.5 Stored energy per unit volume


U 1
u= = ε0 E 2 (4.6)
volume 2

Adapted from Prof. G.H.C. New’s original

5
1st year E & M Key formulae 2 Peter Török

5 Drude model for conduction


5.1 Drift velocity
eE λ
v= (5.1)
2mc
where λ is the mean path between collisions, e and m are the electronic charge and mass, and c is
the mean thermal velocity.

5.2 Current and drift velocity


I = Ne vA (5.2)
where N is the density of free electrons (m−3 ), and A is the cross-sectional area of the conductor.
Note that the direction of I is opposite to that of v although (5.2) deals with magnitudes only.
One can equally write the equation in terms of the current density j (A/m2 ) namely

j ≡ I /A = Ne v

5.3 Conductivity
The conductivity σof a material is defined from the equation

j = σE (5.3)

The resistivity ρ is the reciprocal of σ. According to this simple model of conduction

Ne 2 λ
σ=
2mc
Since E = V /L where V is the potential difference across a conductor and L is its length, it follows
from (5.3) that
L
V =I = IR (5.4)
σA
which is Ohm’s Law. Clearly σis in (Ωm)−1 and ρ is in Ωm.

6 Introduction to Magnetism
6.1 Lorentz Force Law
The force on a particle of charge q moving at velocity v in an electric field E and magnetic field B is

F = q(E + v × B) (6.1)

6.2 Force on a current element


d F = Id s × B (6.2)
where d s is the length in the direction of positive current flow.

6
6.3 Torque on a rectangular current loop in a magnetic field
Γ = IA × B (6.3)
where A is the area of the loop. The equation embodies the sign convention that, looking in the
direction of the vector A, a clockwise current counts as positive.

6.4 Hall Effect


For a conductor carrying current I in a transverse magnetic field B , the Hall voltage is
wIB
VHall = (6.4)
NeA
where w is the width of the conductor (the dimension across which the voltage is registered), A is
its cross-sectional area, and N is the free electron density.

7 Magnetic Field Calculations


7.1 Magnetic flux
!
ΦB = B · dS

7.2 Gauss’s Flux Law for magnetism


"
B · dS = 0 (7.1)

i.e. the flux of B through any closed surface is zero. This amounts to a statement that free magnetic
poles don’t exist.

7.3 Biot-Savart Law (the magnetic field generated by a current element)


µ0 Id s × r̂
dB = (7.2)
4π r 2
where r is the distance from the current element Id s to the point of observation, and r̂ is the unit
vector in this direction.

7.4 Mutual interaction of two current elements


By combining (6.2) and (7.2), the force on element 2 is found to be
µ0 I1 I2 d s2 × (d s1 × r̂)
d 2 F2 = (7.3)
4π r2
where r̂ is a unit vector pointing from 1 to 2.

7.5 Circular loop


At the centre of a loop of radius a
µ0 I
B= (7.4)
2a
where B points in the direction in which the current is seen to flow in a clockwise direction.
On the axis of the loop at a distance z from the centre
µ0 Ia 2
B= (7.5)
2(a 2 + z 2 )3/2

7
7.6 Magnetic field from an infinite straight wire
The radial component of the field is zero. At a distance r from the wire, the azimuthal component is

µ0 I
B= (7.6)
2πr
Looking along the wire in the direction of the current, the field lines circle the wire in a clockwise
direction.

7.7 Force per unit length between two parallel wires carrying currents
µ0 I1 I2
f= (7.7)
2πr
where r is the separation of the wires. If the currents are flowing in the same direction the force is
attractive. The equation is obtained by combining (7.6) and (6.2).

7.8 Ampère’s Law


!
B · d l = µ0 I (7.8)

where I is the current flowing though any surface bounded by the circuit defined by the loop integral.
Current flow is positive in the direction in which the loop integral is seen as clockwise.

7.9 Toroidal coil (on axis)


µ0 NI
B= = µ0 nI (7.9)
2πr
where N is the number of turns and 2πr is the circumference of the toroid; n is the number of turns
per unit length.

7.10 Long solenoid


B = µ0 nI (7.10)
where n is the number of turns per unit length—basically the same result as for a toroid.

Adapted from Prof. G.H.C. New’s original

8
1st year E & M Key formulae 3 Peter Török

8 Electric and Magnetic Fields in Material Media


8.1 Magnetic dipole moment
The magnetic dipole moment of a current loop of vector area A is

m = IA (8.1)

Torque on a magnetic dipole in a field (see Eq.(6.3)) is

Γ = m × B; (8.2)

The potential energy of a magnetic dipole in a B field is

U = −m · B (8.3)

For the electrostatic analogues to Eqs. (8.2) and (8.3), see Eqs. (1.22) and (1.21), respectively.

8.2 Properties of Dielectrics and the Dielectric Constant


In dielectric materials (insulators), electrons are bound to their parent atoms. However, under the
action of an electric field, the charge distribution within an atom is distorted, and an atomic dipole
moment p is induced. The overall effect is to create a large-scale “polarisation” of the medium
defined as P = N p where N is the atomic number density and the polarisation vector P is the dipole
moment per unit volume; note that the dimensions of P are charge per unit area.
If an external electric field E0 is applied to an insulator, the field E within the medium is given by

E = E0 − Einduced

where Einduced is the reverse field arising from the polarisation. Written in vector form, this equation
reads
P = E0
E = E0 − = (8.4)
ε0 εr
where εr is the dielectric constant of the medium.
Note that E0 is traditionally replaced by D/ε0 , where D is known as the electric displacement
vector. The final step in Eq. (8.4) indicates that the dielectric constant is the factor by which the
external field is reduced within the dielectric.
Rearranging Eq. (8.4) yields
D = ε0 E + P = εr ε0 E (8.5)
which is the form normally found in textbooks.

8.3 The Magnetic Analogue


The (albeit imperfect) magnetic analogue to Eq. (8.4) is

B = B0 + µ0 M = µr B0 (8.6)

where B0 is the field from external currents, M is the magnetisation vector, and µr is the relative
magnetic permeability of the medium. The vector B0 is traditionally written µ0 H, where H (in con-
venient units of A/m) is known as the magnetic intensity or the magnetic field, or the H -field (to
distinguish it from B , which some books call the magnetic induction field to distinguish it from H )!!

9
Rearranging Eq. (8.6) now yields

B = µ0 (H + M) = µr µ0 H (8.7)

In free space, µr = 1 and B = µ0 H.


Note that M is small and negative in diamagnetic media, and small and positive in paramagnetic
media; µr ∼ 1 in these cases. In ferromagnetic media, B is large and positive.

Adapted from Prof. G.H.C. New’s original

10
1st year E & M Key formulae 4 Peter Török

9 Time varying currents and electromagnetic induction


9.1 Electromotive Force and EM Induction
9.1.1 Faraday’s Law of Electromagnetic Induction
d ΦB
VEMF = (−) (9.1)
dt
where VEMF is the EMF induced in a circuit, and ΦB is the magnetic flux through the circuit (see
Section 7). The minus sign embodies a sign convention.

9.1.2 Self inductance

The self-inductance of a component (or circuit) carrying current I is defined as

ΦB
L= (9.2)
I
where ΦB is the flux linking the circuit
The EMF induced in an inductor by a changing current is

dI
V i = ±L (9.3)
dt
where the sign is sometimes written + and sometimes - depending on the convention. Using the +
sign brings the formula for the voltage across an inductor into line with the standard formula V = IR
for a resistor.
The self-inductance of a long solenoid is

µ0 AN 2
L= (9.4)
!
where A is the cross-sectional area, N the number of turns, and ! the length of the solenoid.

9.1.3 Mutual Inductance

The mutual inductance of two coils wound on a single solenoid is

µ0 AN1 N2 !
M= = L1 L2 (9.5)
!

Adapted from Prof. G.H.C. New’s original

11
1st$Year$Electricity$and$Magnetism$ Peter$Török$
$
Summary'sheet'1'

1)'Induction'in'conductors'

Electrostatic+induction+in+conductors+is+demonstrated+by+charging+
an+insulating+rod+with+negative+charge+(i.e.+excess+of+electrons)+for+
example+ and+ then+ holding+ it+ close+ to+ a+ conducting+ balloon.+ The+
unbound+ electrons+ in+ the+ conductive+ shell+ of+ the+ balloon+ are+
repelled+ leaving+ positively+ charged+ nuclei,+ which+ are+ attracted+ to+
the+ excess+ negative+ charge+ in+ the+ rod.+ If+ the+ balloon+ and+ the+ rod+
are+allowed+to+touch+the+electrons+from+the+rod+are+transferred+to+
the+ balloon+ which+ then+ will+ have+ overall+ negative+ charge+ thus+
repelling+the+rod.+Note+that+only+about+1+in+1013+electrons+would+
“move+over”+to+the+other+side+of+the+balloon.+

'

2)'Coulomb’s'experiment'

Coulomb+in+1785+found,+by+carrying+out+careful+experiments+
on+ equipment+ we+ now+ call+ Coulomb’s+ torsion+ balance,+ that+
two+ charges,+!! +and+!! ,+ exert+ a+ force+ on+ each+ other.+ The+
forces+ on+ the+ two+ charges+ are+ equal+ in+ magnitude+ but+
opposite+ in+ direction.+ Coulomb’s+ experiments+ showed+ that+
the+ force+ is+ proportional+ to+ the+ charges+ and+ inversely+
proportional+with+the+square+of+the+distance+between+them:+
!! !!
!!" = ! ! )
! ! !"

The+constant+of+proportionality+K+changes+depending+on+what+units+of+measurement+are+used+for+
charge,+distance+and+force.+In+the+SI+system+the+charge+is+measured+in+units+of+Coulombs+[C],+the+
force+ is+ in+ Newtons+ [N]+ and+ the+ distance+ is+ in+ metres+ [m].+ In+ the+ SI+ system+! = 1/4!!! +where+
!! = 8.854×10!!" ![F/m]+is+called+the+vacuum+permittivity+giving+! = 9×10! ![Nm2 /C 2 ].+

3)'The'principle'of'superposition'

The+principle+of+superposition+permits+the+calculation+of+the+
force+ exerted+ on+ a+ single+ charge+ by+ the+ presence+ of+ any+
number+ of+ other+ charges.+ It+ states+ that+ the+ resulting+ force+
can+ be+ calculated+ as+ the+ vector+ sum+ of+ forces+ acting+ on+
charges+pairwise.++

There+ is+ really+ no+ good+ reason+ why+ the+ principle+ of+
superposition+ should+ work.+ It+ is+ simply+ an+ experimental+ fact+
that+ it+ does+ and+ until+ we+ find+ an+ experiment+ that+ proves+
otherwise+ we+ can+ and+ will+ keep+ using+ this+ rather+ useful+
principle.++

+
1st$Year$Electricity$and$Magnetism$ Peter$Török$
$
Summary'sheet'2'

1)'The'electric'field'

The$electric$field$due$to$a$charge$Q$is$defined$as$$

! ! ! N V
! = ! = !!! !!
!!!! C
= m
!
!

where$F$is$the$force$on$a$test$charge$q$placed$at$distance$r$from$Q.$The$electric$field$due$to$many$
charges$ can$ readily$ be$ determined$ from$ the$ principle$ of$ superposition,$ which$ of$ course$ also$
applies$to$the$electric$field$vector.$$

2)'Electric'field'lines'

Electric$field$lines$are$lines$to$which$the$electric$field$
vector$is$tangential$at$any$given$point.$They$are$an$
alternative$ representation$ of$ the$ electric$ field$
instead$ of$ the$ perhaps$ more$ usual$ electric$ field$
vector.$ We$ can$ visualise$ electric$ field$ lines$ by$
floating$ grass$ seeds$ in$ oil$ and$ putting$ an$ electric$
field$ on$ the$ electrodes.$ On$ the$ left$ hand$ side$ the$
grass$seeds$are$disordered$in$the$oil,$but$when$the$
electric$field$is$applied$they$align$themselves$in$the$direction$of$the$electric$field$lines.$

3)'Electric'dipoles'

Electric$ dipoles$ are$ separated$ equal$ and$ opposite$


charges$ and$ they$ can$ be$ created$ for$ example$ by$
separation$ of$ charge$ as$ shown$ in$ the$ figure.$ Electric$
dipoles,$though$not$in$this$form,$are$extremely$common$
in$ nature.$ When$ electric$ dipoles$ are$ placed$ in$ an$
electric$field,$depending$on$their$orientation,$there$will$
be$ a$ torque$ exerted$ on$ them.$ The$ torque$ on$ an$ electric$
dipole$ can$ be$ calculated$ from$ the$ force$ on$ each$ charge$
and$the$distance$from$the$pivot$point:$
! !
Γ= ! sin ! + ! sin ! = !" sin ! = !"# sin !$
2 2
We$ may$ define$! = !!'to$ be$ the$ dipole* moment' with$ a$
being$a$vector$directed$from$the$negative*to*the*positive$
charge.$With$the$definition$we$thus$have$$

! = !×! '

4)'Dipole'moment'of'arbitrary'charge'distributions'

The$ overall$ dipole$ moment$ is$ defined$ as$ the$ vector$ sum$ of$ individual$ dipole$ moments:$! =
! !
!!! !! !i $which$ is$ invariant$ to$ linear$ translation:$ !' = !!! !! (!i + !) = ! + ! !! = ! 'as$
there$are$only$dipoles$present$and$hence$zero$net$charge.$
1st$Year$Electricity$and$Magnetism$ Peter$Török$
$
Summary'sheet'3'

1)'Surface'integrals'

Surface( integrals( are( not( volume( integrals( and( neither( do( they(
result(in(a(volume.(The(result(of(integration(is(either(area,(flux(or(
density(of(function(over(a(given(area.((

Points(to(note:((

a)( Once( the( surface( and( the( function( are( specified( the( integral(
has(a(unique(value(
b)(Integration(can(be(done(in(any(order(
c)( If(! !, ! = 1(then( the( value( of( the( integral( ! ! !, ! !"(is(
the(area(S.(
The( surface( element( in( Cartesian( coBordinate( system( is(!" = !"!# ,( in( spherical( polar( coB
ordinates(is(!" = !! sin ! !"!!"(and(in(cylindrical(polar(coBordinates(it(is(!" = !!!"!!".((

2)'Flux'

Consider(the(surface(S(as(shown(in(the(figure(and(the(2D(vector(field(
V !, ! = ! !, ! ! + ! !, ! !.(Flux(is(defined(as(the(projection(of(the(
vector( field( at$ the$ surface( on( the( surface( normal(!.( The( differential$
flux(at((!, !)(is(given(as(!" = ! ∙ ! !, ! !".(The(total(flux(through(a(
given(surface(is(the(surface(integral(of(the(differential(flux(over(that(
surface:(

!=Φ= ! !, ! ∙ !!!" = ! !, ! ∙ !" (


! !

where(the( ⋯(sign(signifies(that(the(surface(integral(is(carried(out(over(a(closed(surface.(Note(
that(the(flux(is(a(scalar(number(but(it(is(used(to(characterise(vector(functions.(Flux(can(be(positive,(
negative(and(zero(depending(on(the(vector(field(and(the(surface.(Positive(flux(means(that(there(is(
more(vector(field(coming(out(the(volume(defined(by(the(closed(surface(S.(Negative(flux(means(
that( the( vector( field( is( “coming( in”( more( to( the( volume.( Zero( flux( means( a( balanced( in( and(
outflow.(Note(that(it(is(possible(to(define(flux(for(an(open(surface(as(well(as(shown(below.(((

3)'Flux'from'a'point'charge'

The( electric( field( generated( by( a( point( charge( placed( in( the( centre( of( the( surface( S( will( be(
! !
rotationally(symmetric(and(radial:(! = !!! !! .(The(surface(normal(is(also(radial:(! = !.(Thus,(
!

! 1 !
Φ= ! ∙ !" = ! ∙ ! !!" = (
! 4!!! !! ! !!

because(the(surface(integral(integrates(to(surface(of(the(sphere(4!!! .(This(result(is(remarkable(
because( it( shows( that( the( flux( does( not( depend( on( the( radius( of( the( sphere.( Careful(
mathematical(considerations(show(that(the(shape(of(the(surface(does(not(matter(and(the(above(
1st$Year$Electricity$and$Magnetism$ Peter$Török$
$
equation(is(of(very(general(validity.(It(is(also(possible(to(show(that(in(the(case(of(N(charges(the(
result(still(applies:(
!
1
Φ= ! ∙ !" = !! ( (Gauss’s(law)(
! !!
!!!

where( the( summation( is( carried( out( over( all$ charges( within( the( closed( surface( S.( This( result( is(
Gauss’s(law.(This(is(one(of(the(four(fundamental(laws(of(electricity(and(magnetism,(which(means(
that(the(flux(of(electric(field(lines((i.e.(the(electric(flux)(through(closed(surfaces(is(entirely(due(to(
charges(situated(inside(the(surface.(For(if(a(charge(is(placed(outside(the(surface(its(contribution(
to(the(flux(is(zero((what(comes(in(goes(out).(Note(that(Gauss’s(law(is(always$valid(but(its(general(
applicability(depends(on(the(symmetries(in(the(system(under(consideration.(

3)'Electric'field'due'to'an'infinite'charged'plate(

Let( us( determine( the( electric( field( produced( by( a(


charged,( infinite( conducting( sheet( at( a( distance( d(
from( the( plane.( For( Gauss’s( law( to( be( applicable(
the( surface( S( needs( to( be( chosen( carefully.(
Understanding(the(structure(of(the(field(helps(this(
choice.( Since( the( plate( is( infinite,( it( is( plausible(
that( the( electric( field( strength( must( be( the( same(
both(sides(at(equal(distances(from(the(plate.(It(is(also(reasonable(to(assume(that(the(electric(field(
is( perpendicular( to( the( plate.( This( means( that( the( Gaussian( surface( (the( soBcalled( Gaussian$
pillbox)(S(is(chosen(so(that(its(sides(are(perpendicular(to(the(electric(field((i.e.(the(flux(through(
the(sides(is(zero)(and(that(the(top(and(the(bottom(are(perpendicular(to(the(electric(field((i.e.(the(
flux(through(the(top(is(maximum).(The(height(of(the(Gaussian(pillbox(d(is(the(same(above(and(
below(the(plate(so(that(the(flux(through(the(top(and(bottom(is(the(same.(With(these(assumptions(
and(choices(in(mind,(Gauss’s(law(can(be(written(
!
1
! ∙ !" = ! ∙ !" + ! ∙ !" + ! ∙ !" = 2 ! ∙ !" = 2!" = !! (
! top side bottom top !!
!!!

where( A( denotes( the( area( of( the( top( (and( the( bottom)( of( the( Gaussian( pillbox.( This( equation(
becomes(even(simpler(by(noting(that( !! = !"(where(!(is(the(surface(charge(density:(
!" !
2!" = → != (
!! 2!!

This( equation( tells( us( that( the( strength( of( the( electric( field( produced( by( an( infinite( sheet( of(
charges(does(not(depend(on(the(distance(d(from(the(sheet.(This(is(not(an(unexpected(result(of(
course.((

(
1st  Year  Electricity  and  Magnetism   Peter  Török  
 
Summary  sheet  4  

1)  Capacitors  

We   have   seen   for   a   conducting,   infinite   charged   sheet  


that   the   electric   field   strength   was   found   to   be  
! = !/2!! .   Now   we   determine   the   electric   field   due  
to   a   pair   of   conducting,   infinite   sheets,   one   charged  
with   a   surface   charge   density   of  +!  (red),   the   other  
with  – !  (black).   The   electric   field   vectors   due   to   the  
positively   charged   sheet   are   drawn   in   red,   those   due  
to  the  negatively  charged  sheet  are  in  black.  The  black  
field   vectors   are   pointing   towards   the   negatively   charged   plate   because   the   surface   charge  
density  is  negative.  The  principle  of  superposition  permits  us  to  conclude  that  the  field  above  and  
below  the  plates  will  cancel,  whereas  in  between  the  sheets  it  will  double  as  the  vectors  add:  
!
!=  
!!

which   is   the   electric   field   due   to   a   pair   of   charged   conducting   and   infinite   sheets.   Such   an  
arrangement  is  often  referred  to  as  a  capacitor.  Note  that  this  is  only  true  for  infinite  sheets.  In  
case  of  finite  sized  sheets  there  is  a  field  around  the  edges  that  is  often  referred  to  as  the  fringe  
field.    

2)  Line  integrals  

Consider   a   two-­‐dimensional   vector   field   as  


shown  on  the  left  hand  side  of  the  figure  given  
by  !(!, !).  This  may  well  be  thought  of  as  the  
wind   in   an   open   park   whose   direction   and  
strength   changes   from   location   to   location.     If  
we   now   would   need   to   calculate   how   much  
work  W  we  would  have  to  do  against  the  wind  
in  walking  from  A  to  B  it  would  not  be  a  trivial  
task.   It   is   actually   easier   to   calculate   the  
elementary   work   dW   done   (right   inset)   during  
an   elementary   movement   dr   along   the   path  
against   a   force   F,   which   is   given   by   the   dot  
product   of   the   two   vectors,  !" = ! ∙ !! = !! !, ! !" + !! !, ! !"  because  !! = !"! + !"!  
and  ! = !! !, ! ! + !! !, ! !.  The  total  work  W  is  given  as  the  sum  of  the  elementary  work  dW  
done  along  the  path  indicated  by  the  red  line  in  the  figure:  
! !
!= !(!, !) ∙ !! = !! !, ! !" + !! !, ! !"    
! !

Such   integrals   are   called   line   or   path   integrals.   Note   that   the   integral   at   first   glance   looks   just   like  
a   conventional   integral   but   it   is   not   so:   the   relationship   between   the   point   pairs   x   and   y   is   not  
arbitrary  but  determined  by  the  path.  Clearly  the  integral  above  cannot  simply  be  integrated  as  it  
1st  Year  Electricity  and  Magnetism   Peter  Török  
 
contains   both   dx   and   dy.   So   in   solving   these   integrals   the   most   important   part   is   to   first   establish  
this  relationship,  ! !  and  to  substitute  this  relationship  for  y  in  the  integral:  
!!!!
!! !, ! !" + !! !, ! !"   = !! !, ! ! !" + !! !, ! ! ! !(!)    
! !!!!

Once  the  substitution  ! → !(!)  is  done  the  integral  is  no  longer  a  line  integral  but  a  conventional  
one.   Of   course   the  ! → !(!)  substitution   is   equally   valid.   A   three-­‐dimensional   generalisation   of  
the   principle   is   possible.   For   the   sake   of   simplicity   the   path  !  will   now   be   defined   in   the  
parametric  form  ! → (! ! , ! ! , ! ! ):  

!! !, !, ! !" + !! !, !, ! !" + !! !, !, ! !"          


!
!!
= !! ! ! , ! ! , ! ! !(! ! ) + !! !(!), ! ! , !(!) ! !(!) + !! !(!), ! ! , !(!) ! !(!)    
!!

For  some  functions  the  value  of  the  line  integral  depends  on  the  path  but  in  other  cases  it  does  
not.  Those  functions  whose  line  integral  does  not  depend  on  the  path  are  called  conservative  and  
the  corresponding  line  integrals  are  called   path-­‐independent.  It  can  also  be  shown  that  when  line  
integrals   of   conservative   fields   are   calculated   around   a   closed   loop   (the   so-­‐called   loop   integral)  
the  value  of  the  integral  is  always  zero:  

!(!) ∙ !! = 0  
!

where  ! ! = ! !, !, !  is  a  conservative  field.  The  notation   ⋯  signifies  that  the  integration  is  
carried  out  for  a  closed  loop.  Only  for  such  fields  is  it  possible  to  define  a  potential  function.      

3)  Electrostatic  potential  energy  

For  a  single  point  charge  ! ,   electrostatic  potential  energy   U   is  by  definition  the  work  W  that  one  
has  to  carry  out  to  bring  a  charge  !  from  infinity  into  the  vicinity  of  ! :  
! !
!=!= !obs ∙ !! = !el ∙ !!  
! !

where  !obs  is   the   force   experienced   by   the   observer   and  !el  is   the   electrostatic   force   (!el =
−!obs ).  Since  the  charge  is  brought  in  along  a  radial  direction,  the  electrostatic  potential  energy  is  
written  as  
! !
!" !"
!= !el ∙ !! = !
!" =  [J]  
! ! 4!!! ! 4!!! !

Note   that   because   static   electric   fields   are   conservative   it   does   not   actually   matter   which  
direction   the   charge  !  is   brought   in   from.   In   case   of   a   collection   of   charges,   the   electrostatic  
potential  energy  is  calculated  from  the  principle  of  superposition.  

   
1st  Year  Electricity  and  Magnetism   Peter  Török  
 
4)  Electric  potential  

Electric  potential  is,  by  definition,  the  electrostatic  potential  energy  of  a  system  per  unit  charge:  

! J !
!=   =   V  
! C !

which,  for  a  point  charge  is    


! !
!= =  [V]  
! 4!!! !

The   unit   of   electric   potential   is   Volt   [V]   (so   named   after   Alessandro   Volta,   an   Italian   physicist).  
Note  that  the  fact  that  we  can  define  electric  potential  comes  from  the  fact  that  the  electrostatic  
field  is  conservative.    

 
1st  Year  Electricity  and  Magnetism   Peter  Török  
 
Summary  sheet  5  

1)  Field  lines  and  equipotential  surfaces  

The   figure   on   the   right   shows   a  


configuration   of   two   positive   charges,  
one   (!! )   is   charged   to   +4C,   the   other  (!! )  
to   +1C.   Red   lines   are   the   so-­‐called   field  
lines   (field   vectors   are   tangent   to   these  
lines   at   all   points)   and   the   grey   lines   are  
location   of   equipotential   points   (along   a  
single   line   the   electric   potential   is  
constant).  Since  both  charges  are  positive,  
there   is   no   single   location   in   space   where  
the   potential   would   be   negative.   In   the  
far   field   (at   distances   much   larger   than  
the   separation   between   the   two   charges)  
we   may   imagine   that   the   equipotential   lines   are   concentric   circles   looking   identical   to   that  
produced  by  a  single  +5C  charge.  Equipotential  surfaces  come  to  a  point  where  the  electric  field  is  
zero   but   of   course   at   this   point   the   potential   is   non-­‐zero.   This   means   that   one   needs   to   carry   out  
positive  work  to  bring  a  test  charge  to  that  point,  but  once  the  charge  is  there,  there  is  no  net  
force  on  the  test  charge.    

2)  Potential  difference  

We  have  seen  before  that  the  electric  potential  VA  and  VB  at  points  A  and  B  can  be  calculated  as    
! !
!! = ! ∙ !!                              !! = ! ∙ !!.  
!! !!

with   the   potential   of   both   points   referenced   to   infinity.   The   potential   difference   between   the  
points  B  and  A  can  then  be  calculated  as      
! ! !!
∆! = !! − !! = ! ∙ !! − ! ∙ !! = − ! ∙ !!  
!! !! !!

It   may   seem   a   bit   strange   to   write   the   integral   on   the  


right   hand   side   with   the   negative   sign   rather   than  
reversed  integral  limits,  however,  this  is  the  way  most  
textbooks  would  give  this  formula.    

An   upshot   of   this   equation   is   that   if   the   integration  


completes   a   loop,   the   potential   difference   must  
necessarily  be  zero,  hence  

∆! = ! ∙ !! = 0  
!

which  shows  that  the  electrostatic  field  E  is  a  conservative  field.  


3)  Determining  the  electric  field  from  the  potential    

We  have  seen  above  that    


!!
∆! = − ! ∙ !!  
!!

which  implies  that  !" = −! ∙ !!.  But  we  also  know  that    
!" !" !"
!" = !" + !" + !"  
!" !" !"

and  that  ! ∙ !! = !! !" + !! !" + !! !".  Comparing  the  three  previous  expressions  we  have  

!"
!! = −
!"
!"
!! = − → ! = −(!! !! + !! !! + !! !!)  
!"
!"
!! = −
!"
or,  alternatively  

! = −∇! = −grad  !  

Note   that   the   electric   field   in   this   expression   is   a   vector   quantity   and   the   electric   potential   is  
scalar.    
1st  Year  Electricity  and  Magnetism   Peter  Török  
 
Summary  sheet  6  

1)  Ideal  conductors  

In  ideal  conductors  charges  are  allowed  to  move  freely.  This  means  that  (in  the  time  stationary  case)  
there   is   no   electric   field   inside   ideal   conductors   as   when   a   potential   is   applied   charges   will   keep  
moving   until   there   is   no   electric   field   remaining.   All   free   charges   reside   on   the   surface   as   can   be  
shown  from  Gauss’s  theorem.  If  there  is  a  cavity  in  the  conductor,  there  is  no  electric  field  inside  the  
cavity  either.  This  is  the  principle  behind  the  Faraday  cage.    

2)  Surface  charge  density  

Consider   two,   distant   conducting   spheres  


connected  by  a  wire.  One  of  the  spheres,  having  
radius  R1  has  a  charge  Q1,  the  other  with  radius  
R2   has   a   charge   Q2.   Since   the   two   spheres   are  
connected   by   a   wire   they   have   the   same  
potential.  We  calculated  in  class  that  a  sphere  of  radius  R1,  charge  Q1  possesses  a  potential    
!!
!! =  
4!!! !!

and   a   similar   equation   holds   for   the   potential   V2   of   the   other   sphere.   But,   since   they   are  
equipotential   we   have   !! = !!  which   leads   to   !! /!! = !! /!! ,   meaning   that   if,   for   example,  
!! = 10!!  the  charge  on  the  larger  sphere  will  be  10  times  that  on  the  smaller  sphere.  However,  
we   wish   to   investigate   the   charge   density   so   the   areas   of   the   spheres   also   need   to   be   taken   into  
account.  Staying  with  the  previous  example,  !! = 100!! ,  so    
!!
!! = = 10!!  
!!

meaning  that  the  smaller  sphere  has  a  surface  charge  density  10  times  higher  than  that  of  the  larger  
sphere.   Since,   by   Gauss’s   law   the   electric   field   is   directly   proportional   with   the   surface   charge  
density,  the  density  of  the  electric  field  will  be  found  to  be  greater  at  objects  of  smaller  radii.    

3)  Electric  breakdown  and  discharge  

If  an  electron  is  accelerated  in  air,  in  the  presence  of  an  electric  field,  it  can  collide  with  N2  and  O2  
molecules.  If  the  electron’s  kinetic  energy  is  sufficiently  high  it  ionises  the  air  molecules  resulting  in  
an   avalanche   of   electron   production.   When   ions   neutralise   light   is   emitted   which   we   normally  
experience  as  a  spark.  The  voltage  at  which  the  avalanche  process  starts  in  dry  air  is  ~3×10!  V/m  
(also  known  as  the  breakdown  voltage).    

The   knowledge   of   the   breakdown   voltage   permits   one   to   calculate   the   potential   V   and   charge   Q  
stored   by   a   van   der   Graaff   (VDG)   generator.   The   potential   of   the   sphere   on   top   of   the   VDG   is  
! = !/4!!! !  and   so   the   electric   field   on   the   surface   is  ! = !/4!!! !!  giving  ! = !".   Note   that  
this  equation  is  only  valid  for  the  present  geometry.  The  VDG  used  in  the  classroom  has  a  radius  of  
0.15m   so   taking  ! = 3×10!  V/m  we   obtain   for   the   potential   that   the   generator   can   reach   V   =  
450kV  with  a  maximum  charge  of  !!"# = 7.5µμC.  It  is  interesting  to  compare  this  potential  to  that  of  
the   Wimshurst   generator.   The   smaller   and   the   larger   spheres   produce   potentials   of   15   and   30kV,  
respectively.    
1st  Year  Electricity  and  Magnetism   Peter  Török  
 
Summary  sheet  7  

1)  Energy  in  electric  fields  

The  electrostatic  potential  energy  (EPE)  was  defined  as  ! = !!!  which  equals  the  work  it  takes  to  
move   the   charge   q   to   a   point   A,   with   respect   to   which   V   is   calculated.   In   the   past   we   have   looked   at  
what   this   definition   translates   to   in   case   of   two   point   charges.   Now   we   look   at   EPE   stored   in   a  
collection  of  charges.      

Consider  a  collection  of  three  charges,  !! ,  !!  and  !! .  First  we  place  the  first  charge  at  position  1.  As  
there  are  no  other  charges  present  there  is  no  work  done.  The  second  charge  !!  is  now  brought  into  
position  2.  The  work  carried  out  against  the  field  of  charge  1,  i.e.  the  EPE,  is  given  by    

! ! = !! + !! = 0 + !! !!,!  

where  !!,!  means  the  potential  due  to  charge  1  at  position  2.  Then,  moving  charge  3  to  position  3  
one  needs  to  carry  out  work  against  both  charge  1  and  charge  2:  

! ! = !! + !! + !! = 0 + !! !!,! + !! (!!,! + !!,! )  

Consider  now  carrying  the  charges  in  a  reverse  order.  In  this  case  the  EPE  is  given  as  

! ! = !! + !! + !! = 0 + !! !!,! + !! (!!,! + !!,! )  

By  adding  the  two  EPEs  one  obtains  twice  the  EPE  in  the  system:  

2! = ! ! + ! ! = !! !!,! + !!,! + !! !!,! + !!,! + !! (!!,! + !!,! )  

But    !!,! + !!,!  is  just  the  potential  in  position  1,  !! ,  and  so  on.  Therefore  
!
1 1
! = !! !! + !! !! + !! !! = !! !!  
2 2
!!!

with  ! = 3  in  this  case.  In  case  of  a  volume  having  charge  density  !  [C/m! ]  the  elementary  charge  
is   given   as  !" = !"#  with  !"  being   a   volume   element.   Thus   the   total   EPE   stored   in   the   volume  !  is  
given  by:  

1
!= !  !"#  
2 !

where  V  is  the  potential  at  the  volume  element.    

2)  Energy  stored  in  capacitors  

The   new   definition   above   permits   the   direct   calculation   of  


how  much  energy  is  stored  in  a  parallel  plate  capacitor.    

1 ! 1
!= !  !"# = !"# = !"  
2 ! 2 ! 2
But  the  potential  between  the  two  plates  is  just  ! = !"  and  the  total  charge  is  ! = !"  where  !  is  
the  surface  charge  density  and  A  is  the  surface  of  the  capacitor.  Thus  
1 1 1 1
! = !" = !"#$ = !"# = !! ! ! !  
2 2 2 2
which  means  that  the  EPE  per  unit  volume  of  the  capacitor  is  given  by      
! 1
= ! ! !    [J/m! ]  
! 2 !
It   can   be   shown   that   the   expression   is   of   general   validity   with   the   most   general   form   in   vacuum  
being  

1
!= !! ! ! !"  
2 !

3)  Capacitance  

The  capacitance  of  a  capacitor  is  defined  as    

! C !
!=   =   F  
! V !

The  unit  of  capacitance  is  Farad  after  the  great  English  physicist  Michael  Faraday.   The  meaning  of  
the   definition   is   when   the   capacitor   is   charged   how   much   charge   is   needed   to   raise   the   potential   by  
1V.    

The  capacitance  of  a  plane  parallel  capacitor  (see  figure  on  other  side)  can  be  calculated  from  the  
definition  as  follows  
! !" !! !
!= = =  
! !" !
The  energy  stored  in  a  capacitor  in  terms  of  the  capacitance  is  given  by    
1 1
! = !" = !! !  
2 2
1st  Year  Electricity  and  Magnetism   Peter  Török  
 
Summary  sheet  8  

1)  Dielectric  materials  

Electric   fields   applied   to   dielectrics   will  


polarise   materials.   As   shown   in   the   figure  
this   means   that   whilst   the   atoms   in  
dielectrics   have   their   electron   cloud  
distributed  evenly  around  the  nucleus  in  the  
absence   of   an   external   electric   field   (left  
hand  image),  when  an  external  electric  field  
is  applied  (right  hand  image),  the  lower  potential  on  the  left  will  force  the  electrons  to  move  to  the  
right  and  so  the  electron  cloud  will  be  predominantly  located  towards  the  more  positive  potential.    
In  the  case  of  a  capacitor  having  a  dielectric  material  
placed   in   between   the   plates   there   will   be   charges  
!!"#  induced  in  the  dielectric  material  resulting  in  an  
induced   electric   field  !ind .   The   total   electric   field  !  
is   the   vectorial   sum   of   the   electric   field   produced  
without   the   dielectric,  !f ,   and   the   induced   electric  
field   i.e.  ! = !f + !ind  from   which   it   follows   that  
! = !f − !ind .   However,   since  !f = !f /!!  and  !ind = !ind /!!  an   assumption   needs   to   be   made  
regarding   the   fraction   of   the   induced   charge   density.   Assuming   that  !ind = !"f  with  ! < 1  we   may  
write    

1 1
! = 1 − ! !f = !f = !f  
! !!

where  !!  (dimensionless)  is  called  the  relative  permittivity  of  the  material.  Since  for  most  dielectric  
materials  !! > 1,  the  electric  field  E  will  decrease  on  the  introduction  of  dielectrics  in  between  the  
plates.  This  in  turn  leads  to  the  reduction  of  the  potential  V  across  the  capacitor.    

2)  Gauss’s  law  revisited  

The  introduction  of  !!  makes  it  necessary  to  revisit  Gauss’s  law  
!
1
Φ= ! ∙ !" = !!  
! !!
!!!

If   the   medium   the   charges   are   embedded   in   is   dielectric   the   equation   needs   to   be   changed   by  
substituting  !! → !! !! = !  to  give  
! !
1 1
Φ= ! ∙ !" = !! = !!  
! !! !! !
!!! !!!

Note  that  similar  substitutions  need  to  be  made  in  all  those  equations  where  !!  appears.    
1st  Year  Electricity  and  Magnetism   Peter  Török  
 
Summary  sheet  9  

1)  Electric  current  and  current  density  

In  conductors  the  movement  of  charged  particles  (electrons)  causes  current  to  flow.  Such  a  flow  of  
electrons  can  be  caused  by  a  potential  difference.  By  definition,  conventional  current  is  opposite  to  
the  electron  flow.    

The  number  of  electrons  passing  through  a  given  cross  section  of  a  conductor  per  unit  time  is  the  
definition  of  current.  The  current  density  is  defined  as  current  per  unit  area.  

2)  Drift  velocity  

The  velocity  of  free  electrons  in  conductors  is  of  the  order  of  106  m/s.  The  movement  of  these,  so-­‐
called   thermal   electrons   is   random   in   all   directions   within   the   conductor.   When   a   potential  
difference   is   placed   on   two   opposing   ends   of   the   conductors   the   electrons   will   still   move   around  
randomly,  but  overall  there  will  be  a  drift  of  the  electrons  from  the  lower  to  the  higher  potential  end.  
The   velocity   at   which   this   drift   happens   is   called   the   drift   velocity,  !d .   The   drift   velocity   is   many  
orders  of  magnitude  lower  than  the  velocity  of  thermal  electrons,  typically  it  is  of  the  order  of  a  few  
mm/s!   If   for   the   time   being   we   just   assume   that   a   charged   particle   is   drifting   under   the   influence   of  
an   external   electric   field,   the   force   on   that   charge   particle   is   the   Lorentz   force,  !! = !"  with  !  being  
the   charge   of   the   electron.   Under   the   influence   of   this   force   the   particle   will   accelerate   with  
! = !/!! ,  where  !!  is  the  mass  of  an  electron,  to  a  velocity  !d = !"  where  !  is  the  transient  time  
needed  to  reach  drift  velocity.  The  drift  velocity  is  then  given  by  

!"# !"#
!d = =  
!! !! !

where  L  is  the  length  of  the  conductor  on  which  the  electric  field  can  be  expressed  as  ! = !/!.  

3)  The  Drude  model  

Consider   a   conductor   with   N   electrons   per  m! .   There   will   be  !"!d  charge   crossing   a   unit   area   per  
second.  The  amount  of  charge  then  that  passes  through  an  area  !  of  the  conductor  in  1  second  is    

! = !"!d !  

which  of  course  is  the  current.  The  current  density  is  then  given  by  

!
!= = !"!d  
!

or  

!! ! !
!= ! = !"  
!!

where  ! = !! ! !/!!  is   the   conductivity.   So   we   have   arrived   at   the   result   that   the   current   density   is  
simply  given  as  the  product  of  the  conductivity  and  the  electric  field.  

 
4)  Ohm’s  law  

The  current  therefore  can  be  written  as  


!"#
! = !! = !"# =  
!
So  the  potential  is  given  as  

!
!= ! = !"  
!"

which  is  Ohm’s  law.  In  the  above  equation  ! = !/!"  is  called  the  resistance.  The  unit  of  resistance  
is  V/A  or  Ω  (Ohm).  There  are  other  quantities  often  used  to  characterise  materials.  These  are  

! →  conductivity  

! = 1/! →  resistivity  
! !!
! = ! ! = ! ! →  resistance  

Note  that  Ohm’s  law  does  break  down  in  conductors  with  varying  temperature,  as  the  resistance  is  a  
very   strongly   temperature-­‐dependent   quantity.   The   temperature   of   a   conductor   strongly   depends  
on   the   current,   due   to   the   thermal   electrons,   and   so   it   is   seen   that   the   ratio  ! /!  is   no   longer  
constant  as  the  resistivity  also  becomes  a  function  of  the  current,  !(!).    
1st  Year  Electricity  and  Magnetism   Peter  Török  
 
 Summary  sheet  10  

1)  Introduction  to  magnetism  

Magnetism  was  discovered  around  500BC.  The  modern  classical  view  of  magnetism  is  that  it  is  due  
to   current   loops   in   atoms   to   produce   an   overall   magnetic   field.   This   results   in   magnetic   dipole  
moments.   It   is   an   experimental   fact   that   magnetic   monopoles   cannot   be   found.   It   is   possible   to  
artificially   create   materials   that   can   mimic   the   behaviour   of   magnetic   monopoles,   but   they   are  
nevertheless  not  the  real  stuff.  The  easiest  way  to  have  a  handle  on  magnetism  is  by  observing  the  
experimental  fact  that  a  moving  charge  experiences  a  force  in  the  presence  of  magnetic  field.    

2)  Two  experiments  

In  1820  the  Danish  physicist  Hans  Christian  Ørsted  noticed  that  a  compass  
deviates   from   it   stable   position   if   electric   current   flows   through   a   wire  
placed  in  the  vicinity  of  the  compass.  This  was  the  first  known  experiment  
that   connected   electricity   to   magnetism.   If   a   number   of   compasses   are  
placed   around   the   wire   (perpendicular   to   the   plane   of   the   paper)   as  
shown   in   the   figure,   they   will   be   oriented   in   a   tangential   direction   with  
respect  to  a  circle  concentric  with  the  wire  suggesting  that  the  magnetic  
field  induced  in  the  wire  is  also  concentric  with  that.  So  this  experiment  
proves   that   electric   current   exerts   a  
force   on   a   magnet.   It   is   therefore  
reasonable  to  ask  if  it  is  also  true  that  a  
magnetic   field   would   also   exert   a   force  
on   a   current   carrying   wire.   For   this  
purpose   consider   now   a   second  
experiment,  as  shown  in  the  next  figure.  In  the  first  example,  a  permanent  magnet  is  placed  such  
that  it  creates  a  stationary  magnetic  field  across  the  poles.  A  wire  is  then  placed  close  to  the  magnet  
and   current   is   put   though   the   wire   in   such   a   way   that   the   current   flows   into   the   sheet   of   the   paper,  
perpendicular  to  the  magnetic  field.  When  the  current  is  switched  on  the  wire  will  be  pulled  into  the  
magnetic   field.   When,   however   the   direction   of   the   current   is   reversed,   the   wire   will   be   repelled  
away   from   the   magnet.   Thus   we   find   experimentally   that   the   current   I,   magnetic   field   B   and   the  
force  F  exerted  on  the  wire  are  mutually  perpendicular:  

! = !  ×  !  

Note  that  the  experimental  finding  only  implies  the  mutual  orthogonality  of  the  three  vectors  but  
not  the  equation  itself.  The  equation  becomes  valid  because  the  magnetic  field  B  is  defined  via  this  
relationship.    

3)  Force  on  moving  charge  

If  experiments  were  carried  out  on  moving  charges  in  magnetostatic  fields  it  would  be  noted  that  
the  force  acting  upon  a  charge  q  is  given  by    

!! = !(!  ×  !)  

where  v  is  the  velocity  of  the  charge.  It  is  remarkable  that  the  force  only  exists  when  the  charge  is  
moving.  This  finding  forces  an  amendment  of  the  Lorentz  force,  which  is  now  given  by    
! = !! + !! = !! + !(!  ×  !)  

The   unit   of   magnetic   field   is   Tesla  [!"/!"] = [!].   Often   the   non-­‐SI   unit   of   Gauss   is   also   used  
10! ! = 1!.  

4)  Force  on  wire  -­‐  revisited  

Above   we   discussed   how   a   force   is   exerted   on   a   wire  


carrying  charges  if  the  wire  is  immersed  in  an  magnetostatic  
field.   It   is   possible   to   be   a   bit   more   quantitative   about   this  
force  as  shown  next.  Consider  a  section  of  the  wire  as  shown  
in  the  figure.  At  position  A  there  is  a  negative  charge  of  !"  
due   to   a   current   I   flowing   in   the   wire.     The   magnetostatic  
field  is  set  up  as  indicated  and  the  charge  is  moving  with  !!  
drift   velocity   opposite   to   the   direction   of   the   current   flow.  
The  steady  current  flowing  in  wire  is  by  definition  is  given  by  
!"
!=  
!"
and  the  elementary  force  !!  acting  upon  the  charge  by    

!!! = −!" !!  ×  ! = −!  !"   !!  ×  !  


but  since  !! !" = −!!  we  obtain  

!!! = ! !!  ×  !  
which  permits  the  calculation  of  the  total  force  as    

!! = ! !!  ×  !  
wire

where  the  integration  is  carried  out  for  the  wire.  


1st  Year  Electricity  and  Magnetism   Peter  Török  
 
Summary  sheet  11  

1)  Practical  use  of  Lorentz  force  –  particle  accelerators  

The   Lorentz   force,   in   the   absence   of   an   electrostatic   field   is   given   by  


! = !(!  ×  !).  This  means  that  if  a  moving  proton  of  velocity  v  and  charge  
q   is   placed   into   a   magnetostatic   field   whose   direction   is   perpendicular   to  
the   velocity   vector,   the   force   acting   upon   the   proton   will   not   be   able   to  
change  the  kinetic  energy  of  the  proton,  but  it  will  be  able  to  change  the  
direction  of  velocity  as  shown  in  the  figure.  As  a  result  the  proton  will  go  
around  in  a  perfect  circle.  Since  the  velocity  vector,  the  Lorentz  force  and  
the   magnetic   field   are   mutually   orthogonal,   and   the   Lorentz   force   will   be  
equal  to  the  centripetal  force  we  may  write  

!! ! !"
! = !"# = ⇒!=  
! !"

that   is   the   radius   of   the   circle   is   proportional   to   the   momentum   of   the   proton   and   inversely  
proportional   to   its   charge   and   the   magnetic   field.   There   is   an   alternative   approach   to   finding   the  
radius  of  the  circle.  Since  the  kinetic  energy  of  the  of  the  proton  must  be  equal  to  the  electrostatic  
potential  energy  we  have  

1 2!"
!" = !! ! ⇒ ! =  
2 !

that   gives  ! = 2!"/!!! .  Note   that   neither   expression   is   corrected   for   relativistic   changes   that  
can  lead  to  significant  errors  in  calculating  the  velocity  of  the  particle  at  high  potentials.    

2)  Practical  use  of  Lorentz  force  –  Analogue  current  meters  

Consider   a   wire   frame   being   placed   in   a   homogeneous   magnetic  


field  B  with  its  plane  set  parallel  to  the  magnetic  field  direction.  A  
current   I   is   driven   through   the   frame   such   that   it   flows   from  
terminal  A  to  H.    The  Lorentz  force  on  all  segments  apart  from  CD  
and  EF  is  zero  as  these  are  parallel  to  B.  The  Lorentz  force  on  the  
CD  and  EF  sections  is  given  by    
!
!= !(!!×!) = !"#  
!

This  means  that  the  net  force  on  the  frame  is  zero,  but  there  is  a  torque  on  the  frame  !  given  by  
!
!=2 ! = !"#$.  
2
When,   as   a   result   of   the   torque,   the   wire   frame   rotates   to   the   other   extreme   position   where   the  
plane   of   the   frame   is   perpendicular   to   B,   as   the   figure   on   the   right   hand   side   shows,   the   net   force   is  
still   zero,   but   now   the   torque   is   also   zero.     The   practical   use   of   such   a   wire   frame   is   an   analogue  
ammeter   as   it   can   be   seen   that   the   torque   is   proportional   to   the   current   that   is   flowing   through   the  
frame  given  that  the  magnetic  field  and  the  frame  dimensions  are  constant.          
1st  Year  Electricity  and  Magnetism   Peter  Török  
 
Summary  sheet  12  

1)  Biot-­‐Savart  law  

In  1820  Jean-­‐Baptiste  Biot  and  Félix  Savart  discovered  on  a  purely  experimental  basis  
that  if  a  steady  current  I  flows  through  a  wire,  then  the  magnetic  field  around  the  wire  
is  proportional  to  the  current  and  inversely  proportional  to  the  distance  from  the  wire,  
with  the  magnetic  field  direction  looping  around  the  wire  in  concentric  rings:  
!
! ∝  
!
It  is  more  instructive  to  consider  a  small  current  element,  !"!  in  which  case  a  ! !!  dependence  would  
be   expected.   So   the   small   magnetic   field   element  !!  is   given  
by  
!
!! = !   !!×!  
!!
where  C  can  be  experimentally  determined  to  be    
!! Tm !
!= = 10!!   = 10!! !  
4! A !
with  !! = 4!×10!!  [Tm/A]  being   the   magnetic   permeability   of   free   space.   Thus,   the   Biot-­‐Savart  
law  in  differential  form  is  given  by    

!! 1 !! 1
!! = !" !×! = ! !!×!  
4! ! ! 4! ! !

2)  Magnetic  field  due  to  a  current  loop  

Consider   the   geometry   shown   in   the   figure.   The   magnetic  


field  in  the  middle  of  the  loop  is  given  by    

!! 1
!= !! = ! !!×!  
4! !!
!""# !""#

but,  !! ⊥ !  so   !!×! = !".  Therefore  

 
!! 1 !! !
!=! ! !" = !  
4! !! 2!
!""#

because   the   loop   integral   results   in   the   circumference   of   the   loop.  


The  field  lines  due  to  a  wire  loop  look  as  shown  in  the  figure  on  the  
left.   Often   this   field   configuration   is   called   a   magnetic   dipole   field.   It   is   clear   that   no   matter   how   we  
tried  to  put  a  closed  surface  into  this  field  the  flux  would  be  zero,  i.e.  all  vector  field  components  
coming  into  the  surface  eventually  leave  the  surface.  Since  the  flux  is  zero,  we  can  write  
! ∙ !! = 0  
!

which  is  Gauss’s  law  for  magnetic  field  and  is  the  second  Maxwell’s  equation  we  found  in  this  course.  
It  simply  means  that  there  are  no  magnetic  monopoles.  There  is  no  doubt  that  if  they  were  found  
the  Nobel  Prize  would  be  awarded  almost  immediately  for  the  discovery!  

2)  The  Hall  effect  

When  current  flows  though  a  piece  of  conducting  material  


in   the   direction   indicated   in   the   figure,   electrons   move  
with  the  drift  velocity  !!  in  the  opposite  direction.  If  there  
is   a   magnetostatic   field  !  present   the   (magnetic)   Lorentz  
force  !!  will   drive   these   electrons   to   the   bottom   of   the  
slab.   As   a   result   of   the   excess   electrons   on   the   bottom,  
there   will   be   the   (electric)   Lorentz   force  !!  pushing   the  
electrons   towards   the   top   of   the   slab.   When   the   two   forces   balance   each   other   out   there   will   be   no  
net  force  on  the  electrons  which  results  in  a  measurable  voltage  that  turns  out  to  be  proportional  to  
the  magnetic  field:  

!! = !!     ⟹  !" = !!! !      
but  since  ! = !" = !!! !  and  the  current  ! = !"!! ! = !"!! !",  the  potential  is  given  by  

!"
!=  
!"#

Note  that  the  effect  is  sensitive  to  the  orientation  of  the  magnetic  field  and  in  case  of  opposite  !  
direction  the  potential  will  also  change.    
1st  Year  Electricity  and  Magnetism   Peter  Török  
 
Summary  sheet  13  

1)  Ampère’s  law  

We   have   seen   before   that   given   a   current   flowing   in   a   straight   and   infinite  
wire,   as   shown   in   the   figure,   according   to   the   law   of   Biot   and   Savart   the  
magnetic  field  B  around  the  wire  is  tangential  to  a  circle  of  radius  R,  centred  
on  the  wire  and  of  strength  
!! !
!=  
2!"
We  can  choose  to  write  a  loop  integral  for  the  magnetic  field  along  the  circle  to  obtain  

! ∙ !! = !2!"  

because  ! ∥ !!  and   B   is   constant   along   the   circle.   By   virtue   of   the   two   equations   being   equal   we  
thus  obtain  

! ∙ !! = !! !  

which  is  Ampère’s  law  for  time-­‐stationary  currents.  Later  in  the  course  we  will  slightly  modify  this  
law   for   non   time-­‐stationary   cases.   An   important   consequence   of   this   equation   is   that   the   loop  
integral   does   not   have   to   be   done   along   a   circle   because   if  
we   were   to   distort   the   circle   to   an   arbitrary   loop   we   could  
always   resolve   segments   of   the   loop   to   tangential   and  
perpendicular   components.   The   perpendicular   components  
would  not  contribute  to  the  loop  integral  since  there  ! ⊥ !!  .  
Also,  perhaps  even  more  important,  is  the  need  to  define  a  
surface  that  belongs  to  the  loop  as  shown  in  the  figure.  This  
is   necessary   both   to   make   the   definition   of   current   direction  
unique   and   also   to   reveal   which   currents   must   be   taken   into  
account.   As   shown   in   the   figure   both  !!  and  !!  need   to   be  
taken  into  account  as  they  both  penetrate  the  surface  attached  to  the  loop  while  !!  does  not  need  
to  be  considered  as  it  never  crosses  the  surface  assigned  to  the  loop.  If  in  the  figure  the  circulation  
of   the   loop   is   chosen   to   be   clockwise,   the   red   current  !!  is   considered   positive,   while   the   blue  
current  !!  would   be   considered   negative.   If   the   circulation   of   loop   is   defined   in   the   counter  
clockwise  direction  then  !!  and  !!  would  be  considered  negative  and  positive,  respectively.    

2)  Magnetic  field  due  to  solenoids  

Ampère’s   law   can   readily   be   applied   to   solenoids.  


Consider  the  loop  for  integration  as  indicated  in  the  figure.  
The   surface   that   belongs   to   the   loop   is   the   plane   of   the  
paper.   Experimentally   it   can   be   shown   that   solenoids   do  
not   have   magnetic   fields   outside   of   the   coil   and   have  
uniform   magnetic   field   inside.   Consequently,   the  
contribution   to   Ampère’s   law   from   the   top   section   of   the   loop   is   zero.   The   vertical   sides   also   do   not  
contribute   because   they   are   perpendicular   to   the   magnetic   field   inside   the   coil.   Hence   the   line  
integral  gives  

! ∙ !! = !"  

The  right  hand  side  of  Ampère’s  law  needs  to  include  all  current  that  penetrates  the  surface  of  the  
loop.  Hence    
!"!!
!" = !"!! ⇒ ! =  
!
meaning  that  during  the  L  length  of  the  bottom  of  the  loop  the  wires  penetrate  the  surface  n  times.    
Alternatively   one   can   define   the   number   of   turns   per   unit   length  ! = !/!  which   gives   the   final  
result  for  the  magnetic  field  inside  the  solenoid  to  be  

! = !"!!  

3)  Faraday’s  experiment  and  Lenz’s  law  

Michael   Faraday,   after   seeing   Ørsted’s   experiment   with   the   compass,   suggested   that   if   electric  
current  affects  the  compass  then  a  magnetic  field  should  produce  a  current.  In  order  to  prove  this  
he  set  up  two  solenoids,  one  inside  the  other.  He  then  powered  the  one  inside  from  a  battery  and  
noticed  that  there  was  current  induced  in  the  solenoid  outside.  However,  he  has  only  experienced  
current   when   the   switch   was   being   flicked   over.   Once   the   switch  
was   on   the   current   from   the   outside   solenoid   disappeared.   He  
hence   concluded   that   changing   (i.e.   not   steady)   magnetic   fields  
produce   current   in   the   outside   solenoid.   The   phenomenon   is   called  
Electromagnetic   induction.   Lenz   later   experimented   to   find   the  
direction  of  the  current  that  is  produced  by  the  changing  magnetic  
field.   He   found   that   the   induced   current   in   a   current   loop   (shown  
with   arrows   in   the   figure)   is   such   that   its   magnetic   field   (denoted   B   in   the   figure)   opposes   the  
inducing  magnetic  field.  This  is  Lenz’s  law.    

 
1st  Year  Electricity  and  Magnetism   Peter  Török  
 
Summary  sheet  14  

1)  Faraday’s  induction  law  

We  have  seen  previously  that  it  is  the  change  of  magnetic  field  that  results  in  
induction   of   current   in   a   wire   loop.   Consider   a   wire   loop   through   which   we  
have  a  magnetic  field  coming  through  and  it  is  increasing  in  magnitude  over  
time.  By  Lenz’s  law,  the  current  generated  in  the  loop  will  flow  such  that  the  
magnetic   field   it   induces   is   opposed   to   the   incoming   magnetic   field.   The  
current   must   be   produced   by   a   potential   difference   so   there   has   to   be   an  
electric  field  in  the  wire.  The  definition  of  the  electromotive  force  (emf),  ℇ!"# ,  
which  is  the  potential  in  the  wire  is  

! ∙ !! = ℇ!"#  
!""#

Note  that  in  case  of  an  electrostatic  field  the  loop  integral  must  be  zero  as  that  field  is  conservative.  
Electric  fields  induced  by  magnetic  field  are  non-­‐conservative!  

Faraday  carried  out  a  number  of  experiments  with  Helmholtz  coils,  


as   shown   in   the   figure.   He   realised   that   the   induced   emf   in   the  
second   coil,  ℇ!  is   proportional   to   the   change   in   the   magnetic   field  
over   time   produced   by   the   first   coil   and   also   the   area   of   the   second  
coil.   This   permitted   him   to  
conclude   that   the   quantity   of  
importance   is   the   change   in  
the   magnetic   flux   over   time  
passing   through   the   second  
coil.     The       magnetic   flux   is  
defined  in  a  way  similar  to  the  flux  of  electric  field:  

Φ! = open  
! ∙ !!  
surface

Therefore  

!Φ! !
ℇ! = − =− ! ∙ !! = ! ∙ !!  
!" !" open   !"#$
surface !""#

by  virtue  of  the  definition  of  ℇ!  (above).  Note  that  the  definition  of  !!  is  a  matter  of  convention  so  
we  agree  that  the  direction  of  the  circulation  and  that  of  the  surface  normal  is  set  according  to  the  
right  hand  corkscrew  rule.  Also  note  that  the  negative  sign  in  the  above  equation  represents  Lenz’s  
law.  Therefore  we  have  Faraday’s  induction  law  that  is  the  fourth  Maxwell’s  equation  

!Φ! !
= ! ∙ !! = − ! ∙ !!  
!" !"

which  states  that  the  induced  emf  is  due  to  time  changing  magnetic  flux  and  it  opposes  that.    
2)  Some  practical  applications  of  Faraday’s  law  

A   key   to   practical   application   of   Faraday’s   law   is  


to  investigate  the  ways  the  magnetic  flux  can  be  
changed.  For  this  consider  the  wire  loop  shown  
in   the   figure   with   the   current   direction   in   the  
wire  loop  shown  for  increasing  B.  The  magnetic  
flux  can  now  be  written  as    

Φ! = ! ∙ !! = !"#  cos  ! = !"cos  !  

where   A   is   the   total   area   of   the   wire   loop.   Therefore,   there   are   three   possibilities   to   produce   a   time  
varying   magnetic   flux   by   either   changing   the   area   A,   the   angle  !,   or   the   magnetic   field   B:  !"/!",  
!"/!"  and  !"/!".  The  latter  was  discussed  in  relation  to  Faraday’s  experiment  with  the  Helmholtz  
coils,  so  we  shall  look  at  the  former  two.  

Consider   that   the   wire   loop   is   rotated   about   the   axis   shown   by   the   dashed   line   in   the   figure   with   an  
angular   frequency   ! = 2!/!  so   that   ! ! = !" .   The   flux   through   the   surface   is   thus Φ! =
!"cos  !"  so  that  
!Φ!
− = !"#sin  !" = ℇ(!)  
!"
Therefore  the  current  flowing  in  the  circuit  (opposing  the  inducing  magnetic  field  by  Lenz’s  law)  is  
give  by  Ohm’s  law  as  ! = ℇ(!)/!  where  R  is  the  resistance  of  the  loop.  That  is    
!"
! ! = !sin  !"  
!
which  is  called  alternating  current  (AC).  

Next  we  look  at  what  happens  when  the  area  of  the  loop  is  
varied   in   time.   Consider   a   homogeneous   magnetic   field  
with   the   surface   element  !!  upwards   as   shown   in   the  
figure.  According  to  Lenz’s  law  if  the  crossbar  on  the  right  
hand  side  of  the  frame  is  moved  outwards  with  a  velocity  v,  
Φ!  increases   so   the   current   that   is   flowing   in   the   loop   is   in  
a  clockwise  direction.  According  to  Faraday’s  law  we  have    
!Φ! !"
Φ! = ! ∙ !! = !"#     ⟹       = !" = !"# = ℇ(!)  
!" !"
The  current  is  then  given  by  ! = ℇ(!)/! = !"#/!.  It  is  instructive  to  look  at  the  Lorentz  force  on  
the  rod:  

!!! = ! !!×! = ! !!×!       ⟹     !! = !"#  

If  we  want  to  counter  this  force  we  need  to  apply  a  force  of  equal  magnitude  but  opposite  direction.  
Therefore  positive  work  is  done.  Things  do  not  change  if  we  change  the  direction  of  the  movement  
of  the  rod:  the  direction  of  current  changes  direction  so  does  the  Lorentz  force.  Again,  positive  work  
is  done.    
1st  Year  Electricity  and  Magnetism   Peter  Török  
 
Summary  sheet  15  

1)  Displacement  current  

Consider   charging   a   capacitor   as   shown   in   the   figure.   The  


electric  field  is  given  by    
! !free
!= =  
!! !! !!! !! !!

Since  the  current  is  given  as  ! = !"free /!"  the  time  changing  
electric  field  is  given  by    
!" !"free 1 !
= !
= !
 
!" !" !! !! !! !! !! !!

noting  of  course  that  there  is  no  current  flowing  in  between  the  two  plates  of  the  capacitor.  Let’s  
calculate  now  the  magnetic  field  B  at  !!  and    !! .  For  !! :  
!! !
! ∙ !! = !! !     ⇒    ! =  
2!"
For  !!  the  same  calculation  applies  due  to  the  circular  symmetry.  But,  since  there  is  no  current  in  
between  the  plates,  ! = 0,  the  magnetic  field  must  also  be  zero,  ! = 0.    

Consider   now   the   same   capacitor   with   the   open   surface  


belonging  to  !!  defined  such  that  it  is  no  longer  in  the  plane  of  
the  loop  but  the  surface  is  stretched  so  that  it  passes  in  between  
the   plates   of   the   capacitor.   On   this   occasion,   even   though   the  
surface  of  the  loop  is  moved  the  loop  itself  is  not,  we  conclude  
that   the   magnetic   field   must   be   zero   as   there   are   no   current  
lines  penetrating  the  surface.  This  leads  to  the  clear  conclusion  
that  Ampère’s  law  must  either  be  incorrect  or  incomplete.  It  was  
Maxwell   who   set   out   to   solve   this   problem.   The   question   he  
needed   to   answer   is   what   was   so   special   about   the   space   in   between   the   capacitor   plates?   Maxwell  
suspected   that   time   changing   electric   flux   could   possibly   give   rise   to   a   magnetic   field.   Hence   he  
amended  Ampère’s  law:  

!
! ∙ !! = !! ! + !! !! ! ∙ !!  
closed   !" open  
loop surface

where  the  additional  term  is  called  displacement  current.  Let’s  now  apply  the  modified  law  to  the  
problem   of   the   capacitor.   For   the   case   of   the   surface   in   the   plane   of   the   loop,   the   result   is   not  
different   as   the   time   varying   electric   flux   through   the   surface   of   the   loop   is   zero,   so   the   previous  
result   still   holds.   For   the   second   case,   when   the   surface   passes   in   between   the   plates   of   the  
capacitor,  we  may  write  

! ∙ !! = 2!"#  
! = 0  
! ! !" ! ! !
! ∙ !! = !∙! = !! = !
!!! =  
!" !" !" !! !! !! !! !!

Thus  
!! !
2!"# = !! !     ⇒    ! =  
2!"
which  is  the  correct  result.    

The  expression  also  permits  the  calculation  of  the  magnetic  field  inside  the  capacitor  yielding    
!! !  !
!=  
2!!!
 

 
1st  Year  Electricity  and  Magnetism   Peter  Török  
 
Summary  sheet  16  

1)  Magnetic  materials  

In  the  presence  of  strong  electric  fields  electric  dipoles  are  generated  in  materials.  The  overall  effect  
of  these  dipoles  is  the  reduced  effective  electric  field  inside  the  dielectric  material.  The  direction  the  
dipoles  align  depends  on  the  strength  of  the  external  field  and  temperature.    

Very   similar   is   the   case   for   magnetic   materials.   The   presence   of   strong  
magnetic   field   results   in   the   generation   of   magnetic   dipoles   in   the  
material.  If  the  dipoles  are  already  present  these  will  be  aligned  with  the  
external   magnetic   field,   depending   on   the   strength   of   the   external   field  
and   the   temperature.   The   magnetic   dipole   has   an   associated   dipole  
moment  !m  as  shown  in  the  figure.  Magnetic  dipoles  are  due  to  electron  
currents.    

There  are  three  different  types  of  magnetism.  Diamagnetism  is  a  very  weak  effect  that  exists  in  all  
materials.   All   media   respond   to   an   external   magnetic   field   by   producing   a   magnetic   field   that  
opposes  the  field  applied  on  the  material.  It  is  a  similar  phenomenon  to  Lenz’s  law,  but  it  should  not  
be   confused   with   it.   Pure   diamagnets   have   negative   relative   permeability.   We   speak   of  
paramagnetism   when   magnetic   dipoles   already   exist   but   not   in   an   ordered   fashion.   External  
magnetic   field   acts   to   align   these   dipoles   with   the   field.   The   net   Lorentz   force   between   the   external  
field  and  that  of  the  dipoles  results  in  an  attraction  between  paramagnetic  materials  and  permanent  
magnets.   Ferromagnetism   is   characteristic   to   materials   which   have   permanent   magnetic   dipoles  
aligned   in   groups,   or   domains,   within   the   material.   The   presence   of   an   external   field   moves   the  
domain  walls  and  results  in  most  domains  being  aligned  with  the  external  field.  Domain  walls  may  
remain  in  their  new  position  for  a  long  time.  A  mechanical  force  can  however  set  them  back  to  their  
original   position.   Ferromagnetic   materials   lose   their   magnetic   properties   at   high   temperatures  
(typically  many  100s  °C)  when  they  become  paramagnetic.    

2)  Maxwell’s  equations  in  differential  form  

Gauss’s  law  for  electric  fields  in  the  integral  form  reads  
!
! ∙ !! =  
!! !!

The  Divergence  Theorem  also  tells  us  that    

! ∙ !  !" = ! ∙ !  !"  

where   n   is   the   surface   normal   of   the   closed   surface   over   which   the   left   hand   side   surface   integral   is  
carried  out.  Hence  

! ∙ !! = ! ∙ !  !"  

but  ! = ! !"  so  


1 !
! !" = ! ∙ !  !"     ⇒  ! ∙ ! =                                                                                                ME1  
!! !! !! !!
   

Gauss’s  law  for  magnetic  fields  is  given  by  

! ∙ !! = 0  

and  so  the  direct  use  of  the  Divergence  Theorem  yields  

 ! ∙ ! =  0                                                                                                                                                          ME2  

Ampère’s  law  for  magnetic  fields  is  given  by  


!
! ∙ !! = !! !! ! + !! !! ! ∙ !!  
!"
Stokes’  Theorem  states  

! ∙ !! = !×! ∙ !!  

and    ! = ! ∙ !!  with  j  being  the  current  density.  We  therefore  have  


!
!×! ∙ !! = !! !! ! ∙ !! + !! !! ! ∙ !!  
!"
from  which  we  have  
!!
!×! = !! !!  ! + !! !! !! !!  
!"
but  since  !! !! = ! !!  and  !! !! = !! ,  the  above  equation  can  be  written  

!! !!
!×! = !! !!  ! + !                                                                                                                              ME3  
! !"

Faraday’s  induction  law  is  given  by    


!
! ∙ !! = ! ∙ !!  
!"
since,  by  Stokes’  theorem   ! ∙ !! = !×! ∙ !!,    

! !!
!×! ∙ !! = ! ∙ !!             ⇒       !×! =                                                                        ME4  
!" !"

 
1st year E & M Electrostatic force, electric field, and dipoles Problem sheet - Week 1

Useful constants and information:

12
• "0 = 8.854 ⇥ 10 F m 1; • g = 9.807 m s 1 ;
19
• e = 1.602 ⇥ 10 C;
• Atomic number density of Cu = 8.5 ⇥
31
• me = 9.109 ⇥ 10 kg; 1028 m 3 .

Some simple exercises to get you going


(Answers are given at the end)

A. Calculate the ratio of the electrostatic to the gravitational force between two electrons. Why is
the ratio independent of their separation?

B. Consider two 1 cm3 blocks of copper 10 m apart. If, in each block, one electron is removed
from one atom in a million, find the magnitude of the repulsive force between the blocks.
Roughly how many people could the force support?

C. An electric dipole consists of charges ±2.0 ⇥ 10 8 C separated by 1 cm. What is the magnitude
of the electric field at the centre of the dipole?

Problems
Starred questions are a bit more challenging than the unstarred ones.

1. A dipole is formed along the x-axis by a charge +Q at x = a /2 and a charge Q at x =


+a /2. Show that the electric field at a point C = (0, y) ⌘ (0, (a /2) tan ↵) along the y-axis is
(2Q cos3 ↵)/(⇡✏o a 2 )î

2. An electric dipole is formed by two charges +Q1 and Q1 and situated on the x-axis at
x = a /2 and x = +a /2 respectively. A second dipole is formed by two further charges
of magnitude Q2 , also on the x-axis at x = X ± b /2. The dipoles are widely separated (so that
X a, b). If the second dipole is in its stable (lower energy) alignment with respect to the
first, draw a diagram showing:

(a) the direction of the E field of the first dipole in the vicinity of the second;
(b) the positions of all four charges;
(c) the directions of the two dipole moment vectors.

3. ***An electron is released from rest in a uniform electric field. The electron accelerates verti-
cally upward, traveling 4.50 m in the first 3.00 µs after it is released. (a) What are the magni-
tude and direction of the electric field? (b) Are we justified in ignoring the effects of gravity?
Justify your answer quantitatively.

4. *** Consider a particle of mass m and charge +Q situated directly beneath a dipole as shown
in the diagram, top of next page. If the three charges have equal magnitudes of 3 nC, and
are equally spaced 1 cm apart, find the value of m for which the particle is held in equilibrium
under gravity.
Is the equilibrium stable or unstable?
3. Consider a particle of mass
+Q beneath a dipole as shown in the diagram. If the thr
have equal magnitudes of 3 nC, and are equally
find the value of
−Q under gravity.

Is the equilibrium stable or unstable?


m +Q

5. Assessed Problem
4.It was shown in the lectures that the mutual potential energy of three charges
(a) An electric dipole is formed byQtwo
3 charges of ±1nC spaced 4 mm apart. Find the exact
magnitude of the electric field on the axis Q
of theQdipole
Q at 18 mm from the centre (see
diagram) + [1 mark]

4 mm r12

x
Q V +Q
-1 nC +1 nC
Vi
18 mm

(b) Compare your answer to part (a) with E ⇡ p /2⇡✏0 x 3 which is a good approximation at
large x, and express the result as a fraction of the exact result. [1 mark]

Answers
A 4.16 ⇥ 1042 . 2 -
B 26 people (assuming an average weight of 3 (a) |E| = 5.69 N/C, downward; (b) -
65kg = 143lb = just over 10 stone).
C E = ↵(i + j); (ii) E = (yzi + zxj + xyk) 4 6.18 ⇥ 10 5 kg. -

1 - 5 -

Peter Török
st
Solutions – Problem sheet – week 1 – 1 year E&M

! e2 $ ! Gm 2 $ e2
#" r 2 &% = 4!" Gm 2 = 4.16 ' 10
42
A. # 2&
" 4!" 0r % 0

The result is independent of r because both the electrostatic and the gravitational forces obey an
inverse square law.

B. Charge in each block Q = 1.6 ! 10"19 ! 8.5 ! 1028 ! 10"6 ! 10"6 = 1.36 ! 10"2 C;

Q2
= 1.66 ! 104 N
4!" 0102

A 65 kg (=10 st 3 lb) person weights 637 N, so roughly 26 people of this weight could be
supported.
C.
2Q
E= = 1.44 ! 107 V/m
4!" 0 (d / 2)2

1.

2.

3. Uniform E. Only solving in vertical (y) direction, up -> +.


!!! ! !"
!! = −! ! = !! with ! = 0 = !"/!" at ! = 0. Simple acceleration with solution ! = − ! !.
!! ! ! !!
Substituting quantities given in question and solving for E results ! = −5.69N/C (downward). (b).
Gravity can be ignored when the graviational force is much smaller than the electric force. The ratio of
gravitational force and electric force is 9.81×10!!" so the graviational force can be ignored.

Q2 Q2 3Q 2
4. Upward force on mass = ! = = mg to balance.
4!" 0 a 2 4!" 0 (2a) 2 16!" 0 a 2
3Q 2 3 ! 9 !10"18
m= = "12 "4
= 6.18 !10"5 kg . Unstable
16!" 0 a g 16! ! 8.85 !10 !10 ! 9.81
2

equilibrium.
1st year E & M Gauss’s law, electric field and flux Problem sheet - Week 2

Useful constants and information:

• G = 6.673 × 10−11 N m2 kg−2 ; • me = 9.109 × 10−31 kg;

• ε0 = 8.854 × 10−12 F m−1 ; • g = 9.807 m s−1 ;

• e = 1.602 × 10−19 C; • a0 = 5.29 × 10−11 m

Some simple exercises to get you going


(Answers are given at the end)

A. Find the electric flux through a closed surface surrounding the charges -2.5µC, -1.3µC, +0.8µC
and +3.2µC.

B. The charge density (Coulombs per cubic metre) within a sphere of radius a varies as ρ(r) =
β(r /a)n where β is a constant, n is an integer, and r is distance from the centre. Show that the
total charge in the sphere is Q = (4πβa 3 )/(n + 3)

C. A copper plate 1 mm thick is placed in a strong external electric field Eext = 106 V/m normal to
the surface. If copper has 8.5 × 1028 free electrons per m3 , what proportion of these migrate
to the surface to maintain zero internal field? Might there be a shortage of electrons as the
plate is made thinner?

Problems
1. Using your lecture notes as little as possible (preferably not at all!) and Gauss’s Flux Law in
every case, reproduce the following electric field calculations done in the lectures:

(a) Show that the electric field of a uniformly charged spherical shell of radius a carrying
total charge Q is
Q
(r > a)

πε
 2
4 0r



E=


(r < a)

 0

(b) Show that the electric field at a distance x from a infinite line charge of linear charge
density λC/m is
λ
E=
2πε0 x
(c) Show that the electric field adjacent to a infinite charge sheet of areal charge density σ
C/m2 is
σ
E=
2ε0

2. The charge density on a plate, due to a charge Q placed at a height h above the point A on
the plate, is
−Qh
σ(r) =
2π(r + h 2 )3/2
2

where r is the distance from the point A on the surface of the plate. Find the total charge on
the plate that lies within r = 10cm of A if the charge Q = 50nC us placed at h = 1cm.
3. In the lectures, Gauss’s Flux Law was used to show that the electric field a distance x from
an infinite line charge lying along the y-axis is

λ
E(x) =
2πε0 x

where λ (C/m) is the charge per unit length.


Now obtain this answer by direct integration. This is not only a good exercise; if you want to
find the field from a line charge of finite length, you have to do it this way!
The question leads you through the calculation in easy stages. Start by finding the contribution
from an element dy and go on to integrate over all such elements.

(a) Show that the magnitude of the contribution


to the electric field at P (see diagram) from the
dy
2
cos θ dy

y
of the field from
responding
−y θ
P
cos 3 θ dy x

(c) The problem can now be solved by integrating either


1
θ 2 π
(a) Show that the magnitude of the contribution to the electric field at P (see diagram) from
the element
xdθ dy at y is
λ cos2 θ dy
dE =
4πε0 x 2
(b) Show that the x-component of the field from this element, taken together with the corre-
sponding element at −y is
λ cos3 θ dy
dEx =
2πε0 x 2
(c) The problem can now be solved by integrating either over θ from 0 to π/2, or over y from
0 to ∞. To take the former route, first show that dy = (xd θ)/ cos2 θ
(d) Perform the integration
ε
to obtain the result.
(e) If you’re on a roll,ε tryε εthe other integration option suggested in (3c).
This standard integral will come in handy:
Z
dy y
/
= 2 2
(a 2 2
+y ) 3 2 a (a + y 2 )1/2

4. *** Consider a thin circular disk of radius a lying in the x − y plane with its centre at the origin
and carrying a uniform charge density σ. Find the z-component of the field from a thin annular
ring directly, and then integrate over the disk. Check for the special case of α → π/2 (α being
the angle at which the rim of the disk is seen from the observation point on axis) the result
yields the electric field of an infinite sheet. Now check what happens when α → 0. Why does
this limit not yield the electric field due to a point charge. How can you rectify the problem?

5. Assessed problem Consider a uniformly charged sphere of radius a carrying charge density
ρ from which a spherical section of radius b (< a) has been removed to leave a hollow cavity.
The centre of the cavity is at d (< a − b) from the centre of the larger sphere (see diagram).
y

d
a
b x

Calculate the electric field


Show that everywhere
the electric field is E ( ρ within the cavity.
/ 3ε 0 )i everywhere within the cavity. [4 marks]

Answers to Exercises
A 2.26 × 104 Vm; 2 −45nC

B - 3 (d) and (e): λ(2π0 x)−1


 
σz −1 1
C 6.5 × 10−13 λ 4 Ez = 20 (a 2 +z 2 )1/2
+ z
λ
1 - 5 -

Peter Török
st
Solutions – Problem sheet – week 2 – 1 year E&M

1. In notes
2.
⎛ −Qh ⎞
dQ = ⎜ 2 3/ 2 ⎟
2π rdr
⎝ 2π (r + h ) ⎠
2

R
R
rdr ⎡ 1 ⎤ ⎡ 1 1⎤
QR = −Qh ∫ 2 = −Qh ⎢− 2 2 1/ 2 ⎥
= −Qh ⎢− 2 + ⎥
0 (r + h ) ⎣ (r + h ) ⎦ 0 ⎣ (R + h )
2 3/ 2 2 1/ 2
h⎦
⎡ 1 ⎤
QR = −Q ⎢1 − 2 1/ 2 ⎥
⎣ (1 + ( R / h) ) ⎦
" 1 %
QR = !Q $1! ' = !45.0nC
# 101 &
λdy λ cos2 θ dy
= =
3. (a) Field at P due to dy element
4πε 0 r 2 4πε 0 x 2 since x = r cosθ .
(b) Multiply by a further cos θ to resolve the field in the x-direction, and double the result to
include the corresponding element at -y.

(c) y = x tan θ so dy = x sec2 θ dθ .

π /2
λ cosθ dθ λ
[sin θ ]0 = λ
π /2
∫ 2πε 0 x
=
2πε 0 x 2πε 0 x
(d) Total field = 0
x
cosθ =
(e) OR set
x 2 + y 2 to obtain
Total field
∞ ∞
λx 3 dy λx ⎡ y ⎤ λx ⎡ 1 ⎤ λ
=

2πε 0 x ( x + y )
0
2 2 2 3/ 2
=
2πε 0
⎢ ⎥ =
⎢⎣ x 2 x 2 + y 2 ⎥⎦
0

2πε 0 ⎣ x 2
− 0⎥ =
⎦ 2πε 0 x
as before.

(f) If you try to work out the potential, you get an infinite result. The potential is related to the work
done in bringing a charge from infinity up to the wire, and this really is infinity for a wire of infinite
length. All real wires are of course of finite length.
4).
a
2! r" cos # " "z # "1 & "z # "1 1&
a a
rz
Ez = ! dr = ! dr = % 2 1/2 (
= % + (
0
4!$ 0 (r + z )
2 2
2$ 0 0 (r + z )
2 2 3/2
2$ 0 $ (r + z ) ' 0 2$ 0 $ (a + z )
2 2 2 1/2
z'
z
cos ! = 2 2
where r + z resolves the field in the x direction. It follows that

! " z % !
Ez = $1! 2 2 ' =
2" 0 #
[1! cos# ]
a + z & 2" 0
for ! ! " / 2 the equation trivially returns the value for an infinite sheet. For ! !0 it seems to
give the wrong answer. We can re-write the result as

Q
Ez =
2!" 0 a 2
[1! cos! ]
where a = z tan ! with z being the axial distance between disk and observation point. If this
expression is substituted back to the formula for Ez and the L’Hospital rule is used, one obtains
the equation for a point charge.
1st year E & M Electrostatic potential energy, electric potential Problem sheet - Week 3

Useful constants and information:

• G = 6.673 × 10−11 N m2 kg−2 ; • me = 9.109 × 10−31 kg;

• ε0 = 8.854 × 10−12 F m−1 ; • g = 9.807 m s−1 ;

• e = 1.602 × 10−19 C; • a0 = 5.29 × 10−11 m

Some simple exercises to get you going


(Answers are given at the end)

A. Find the electric field if the electrostatic potential is (a) V = α(x + y) and (b) V = βxyz, where
α and β are constants.

B. Sketch the electric field lines from charges -Q and +2Q a distance a apart. Consider the
following points:

(a) Will the field pattern be symmetric?


(b) What does the field pattern look like very close to each charge?
(c) At what distance from the −Q charge is the electric field zero?
(d) What does the field pattern look like a long distance ( a) from the charges?

You might want to look at the field line configuration using one of the JAVA applets provided
on the course website.

Problems
1. Find the magnitude of the electric field at a distance of one Bohr radius (a0 ) from a proton; this
is the field experienced by an electron in the first Bohr orbit (the ground state) of a hydrogen
atom.
Calculate the potential energy of the electron and proton taking the zero of potential energy
to be at infinite separation. Express your answer (a) in Joules and (b) in electron volts (eV).
Now assume the electron orbits the proton in a circular path of radius a0 . Find (a) the orbital
velocity of the electron, (b) the ratio of potential to kinetic energy, and (c) the total energy of
the electron (= the ground state energy of hydrogen).
[Ignore any radiation of energy associated with the electron’s acceleration. One of the Bohr
postulates is precisely that there isn’t any!]

2. Charges are located at each of the four corners of a square of side a. Find the potential
energy for each of the following configurations:

(a) Four identical charges +Q;


(b) Two charges +Q and two charges -Q, with charges of like sign at opposite corners;
(c) Two charges +Q and two charges -Q, with charges of like sign at adjacent corners.

Could the charges be described as bound by the electrostatic forces in any of the three cases
and, if so, in which?
3. In connection with the energy stored in a capacitor, we showed in lectures that the mutual
potential energy of three charges Q1 , Q2 and Q3 is
!
1 Q1 Q2 Q1 Q3 Q2 Q3
U= + +
4πε0 r12 r13 r23

where r12 is the separation of Q1 and Q2 etc. Show that the result may be written in the form

1
U= (Q1 V1 + Q2 V2 + Q3 V3 )
2
where Vi is the potential at Qi due to the other two charges.

4. Show that, in the far field approximation (x  d), the electric potential of a dipole formed by
two charges ±Q situated at (±d /2, 0, 0) on this x axis is given by
p
V=
4πε0 x 2

where the dipole moment p = Qd. Obtain the associated electric field by differentiation.

5. Obtain expressions in the far field approximation (r  d) for the electric potential and electric
field at a point P, distance r (measured from the mid-point of the dipole) away from a dipole
formed by two charges ±Q situated at (±d /2, 0, 0) (see figure below).

P(x,y)

r+

r r-

ϑ
d/2 d/2
You may like to follow steps (a) to (c) initially:

(a) Obtain an approximate expression for r+ and r− as a function of r.


(b) Use an exact formula to obtain the electric potential due to both charges at P with dis-
tances r+ and r− .
(c) Use a series expansion on the distances assuming that r  d to obtain
px
V
4πε0 r 3

(d) Obtain Ex by differentiating V and not forgetting that r = f (x):

∂V p
Ex = − = (3 cos2 θ − 1)
∂x 4πε0 r 3
where θ is the angle between the mid-point and the x-axis (see figure).
(e) Obtain the corresponding expressions for Ey and Ez .
(f) Show that Ey = Ez = 0 on the y-axis.
(g) Show that the locus√ of points for which Ex = 0 in the far field is a cone whose vertex
angle is 2 cos−1 (1/ 3).
6. *** An electric dipole is formed by two charges +Q1 and −Q1 and situated on the x-axis at
x = −a /2 and x = +a /2 respectively. A second dipole is formed by two further charges of
magnitude Q2 , also on the x-axis at x = X ± b /2. The dipoles are widely separated (so that
X  a, b). Obtain expression for the electric field on the x axis at distances far away from
the dipoles (“far-field approximation”). From this result obtain the potential energy difference
between the opposite orientations of the second dipole in the field of the first dipole.
7. ***The so-called van der Waals force between neutral atoms and molecules arises from a
“dipole-dipole” interaction. The incessant movement of the bound electrons means that an
atom or molecule will exhibit a fluctuating dipole moment p1 , which in turn creates an induced
dipole moment p2 in a neighbour.
Making the reasonable assumption that the magnitude of p2 is proportional to the field of
p1 , convince yourself that the van der Waals force varies as the inverse 7th power of the
separation between them (i.e. force ∼ 1/r 7 where r is the distance between the molecules).
You may want to use results in Question 6 above. Is the van der Waals force attractive or
repulsive?
8. Assessed problem
(a) A small sphere of radius a carries charge q distributed uniformly over the surface. From
the formula for the capacitance and the expression for the energy stored in a capacitor,
show that the stored energy is
q2
U=
8πε0 a
[2 marks]
(b) Obtain the same result by working out the total energy stored in the electrostatic field.
Use the fact that the energy density (J/m2 ) in an electrostatic field in vacuo is (ε0 E 2 )/2.
[1 mark]
(c) The electron appears to behave in most respects as a point particle. However, if one
assumes that it is actually a small sphere, one can estimate its radius by equating the
stored electrical energy to its rest energy mc 2 . Show that this predicts an electron radius
of
e2
R= .
8πε0 mc 2
Put in the numbers to calculate the value of R.
[2 marks]

Answers
A (a) E = −α(î + ĵ), (b) E = −β(yz î + xz ĵ + xy k̂) 4 Ex = p /(2π0 x 3 )
B -
5 -
1 −2.17 × 10−18 J.
√ √ 6 E ≈ Q1 a î/(2π0 x 3 ), ∆U ≈ Q1 Q2 ab /(π0 X 3 )
2 (a) V = Q 2 ( 2 + 4)/(4π0 a), (b) V√= Q 2 ( 2 −
4)/(4π0 a), (c) V = Q 2 (− 2)/(4π0 a).
7 -, attractive
Bound in (b) and (c)
3 - 8 -

Peter Török
st
Solutions – Problem sheet – week 3 – 1 year E&M

A.

B.

1.
3.

4.

5. a-c
6.

7. Since

and

According to the expression for E, the field of p1 varies as X-3, hence !! ~! !! and from the result for F
above !~!! ! !! = ! !! . The van der Waals force is attractive.
1st year E & M Gauss’s law, electric potential, capacitance Problem sheet - Week 4

Useful constants and information:

• Capacitance C = Q /V ; • e = 1.602 × 10−19 C;

• ε0 = 8.854 × 10−12 F m−1 ; • Stored energy U = 12 Q 2 /C ≡ 12 CV 2

Some simple exercises to get you going


(Answers are given at the end)

A. The plates of a parallel-plate capacitor (area 8 × 10−3 m2 ) are separated by a 0.5 mm air gap.
Assuming that the dielectric constant of air εair
r  1, calculate the capacitance.

B. If the charge on the plates is ±5 nC, find the potential difference V between the plates; the
electric field E between the plates; and the stored energy U in the capacitor.

C. For a capacitor that is isolated (i.e. the charge is constant), consider what happens when the
plate separation is doubled. What is the effect on the capacitance C; the potential difference
V; the electric field E; and the stored energy U?

Problems
1. A spherical insulator of radius a carries uniform charge density ρ and total charge Q. It is
easy to show from Gauss’s Law that the electric field outside the sphere is given by
Q
E(r) = (r > a) (1)
4πε0 r 2
Show that inside the charge distribution (r < a), the field is
Qr
E(r) = (r < a) (2)
4πε0 a 3

2. It was shown in the lectures (both theoretically and experimentally) that charged spherical
conducting shells have zero electric field inside. It is an interesting problem then what happens
when we put a charge inside without touching the surface of the shell.

a
+Q
b

(r > b) r < b)

So consider the case of a conducting sphere of radius b containing a concentric spherical


cavity of radius a(< b), and a point charge +Q at the centre (see diagram). Assuming there is
no net charge on the conductor, find the electric field (a) outside the sphere (r > b) (b) within
the conductor (a < r < b) and (c) in the central cavity (r < a).
What is the total charge on the inner and outer surfaces of the conductor?
3. By integrating Eqs. (1) and (2) in Question 1, show that the corresponding electric potentials
are
Q
V= (r > a)
4πε0 r
r2
!
Q
V= 3− 2 (r < a)
8πε0 a a

Draw rough graphs of E(r) and V (r). Is ∂V /∂r continuous at r = a?

4. For the case of Question 2, integrate your expressions for E(r) to obtain corresponding ex-
pressions for V (r). Ensure that the potential is continuous at the boundaries. As in Question
3, draw rough graphs of E(r) and V (r). Try to understand the graphs physically.

5. *** A capacitor is formed by two coaxial conducting cylinders of length `. The radii of the inner
and outer cylinders are a and b respectively.

(a) The outer and inner conductors carry linear charge densities +λ and −λ (C/m) (or equiv-
alently surface charge densities λ/2πb and λ/2πa). Find an expressions for the capaci-
tance of the pair of cylinders.
(b) If ` = 100mm, a = 10mm and b = 11mm, find the capacitance. Find also the capacitance
of a plane parallel capacitor having the same plate area as that of the inner cylinder, and
the same 1 mm plate separation.

6. Assessed problem Charge is distributed uniformly (charge density σ) over the surface of
a flat circular disk with a central hole (see diagram). The inner and outer radii are a and b
respectively as shown.

(a) Obtain an expression for the electric potential at the centre of the disk (O).
[1 mark]
(b) Obtain an expression for the electric potential at P, on the axis of the disk at distance z
from the centre.
[1 mark]
(c) Hence show that the electric field at P is

σz
!
1 1
Ez = √ − √
2ε0 a2 + z2 b 2 + z2
[1 mark]
(d) Show that for small z ( a, b)
σz b − a
Ez ≈
2ε0 ab
[1 mark]
(e) A charged particle at O may be stable or unstable to small displacements along the z-
axis depending on the sign of its charge. For the stable case, obtain an expression for
the period of small oscillations of the particle as a function of its charge and mass.
[2 marks]

Answers to Exercises
A 142pF a<r <b

B 35.3V, 70.6kV/m, 88.2nJ 3 -

C /2, ×2, no change, ×2 4 -

1 - 5 C = 2π0 `/(log(b /a)), 55.61pF

2 E = Q /(4π0 r 2 ) for r > b and r < a; E = 0 for 6 -

Peter Török
st
Solutions – Problem sheet – week 4 – 1 year E&M

1. Applying Gauss’s Flux Law to a spherical Gaussian surface of radius r (< a)


4πr 3ρ ρr Qr 4πa 3 ρ
E × 4πr =
2
which leads to E = = since = Q.
3ε 0 3ε 0 4πε 0 a 3 3

Q
2. (i) E (r ) = ( r > b)
4πε 0 r 2
(ii) E (r ) = 0 ( a < r < b)
Q
(iii) E (r ) = (r < a)
4πε 0 r 2

Charge + Q resides on the outer surface and − Q on the inner surface.

Q
3. V (r ) = (r > a)
4πε 0 r
Qr 2
By integrating eq.(2), V (r ) = − + K , where K must be chosen to
8πε 0 a 3
Q
ensure continuity of V at r = a , i.e. that V (a ) = . This yields
4πε 0 a
3Q
K= and the result follows immediately.
8πε 0 a

∂V
E (r ) = − is continuous at r = a .
∂r

Q
4. V (r ) = ( r > b)
4πε 0 r
Q
V (r ) = ( a < r < b)
4πε 0 b
Q
V (r ) = +K (r < a)
4πε 0 r
where K must be chosen to ensure continuity of V at r = a . This leads to
Q ⎛1 1⎞
K= ⎜ − ⎟ and it follows that
4πε 0 ⎝ b a ⎠

Q ⎛1 1 1⎞
V (r ) = ⎜ + − ⎟
4πε 0 ⎝ r b a ⎠

5. (a) Using Gauss’s Flux Theorem


λl λ
E × 2π r l = yields E = (pointing radially outwards)
ε0 2πε 0 r


Vouter − Vinner = − Edr
b
λ λ
Vinner − Vouter =

dr
= [log r ]ba = λ log b
2πε 0 r 2πε 0 2πε 0 a
a
Q 2πε 0
C= =
V log(b / a)

2πε 0 0.1
(b) C = = 58.34 pF
log(1.1)

For the plane parallel case


ε 0 2π al
C= = 55.61pF
(b − a)
1st year E & M Resistivity, Lorentz force Problem sheet - Week 5

Useful constants and information:

• F = q(E + v × B); • mp = 1.67 × 10−27 kg

• Velocity of light = 3 × 108 m/s • me = 9.11 × 10−31 kg

• Atomic number density of Cu = 8.5 ×


• µ0 = 4π × 10−7 H m−1 ; 1028 m−3 .

• e = 1.602 × 10−19 C • The resistivity of copper is 1.7 × 10−8 Ωm.

Some simple exercises to get you going


(Answers are given at the end)

A. A piece of copper wire radius 0.5 mm carries a uniformly distributed current of 1 A. Find the
electron drift speed in the wire.
[For comparison, the thermal speed of the electrons at 20◦ C is of the order of 105 m/s.]

B. If the wire in Part A above is 100 m long, calculate the energy dissipated in the wire in one
minute.

C. A straight copper bar with rectangular cross section is aligned along the z-axes and carries a
current of 3A in this direction. A magnetic field B of 1 T is applied in the x-direction, perpen-
dicular to two sides of the bar that are separated by 1.5 mm. In what direction does the Hall
effect voltage appear, and what is its magnitude?

Problems
The electron was discovered over 100 years ago by J.J. Thomson working at the Cavendish Labora-
tory in Cambridge. The details contained in this Classwork are not an exact reflection of Thomson’s
experiments between 1897 and 1899, but they are close, and the concepts are the same. Problems
1-4 are all more challenging than usual unstarred questions.

1. A beam of electrons moving at a controlled speed pass through a uniform electric field created
between two parallel plates of length L , separation d, and potential difference V . As the
electrons enter the region between the plates, they are moving at speed v0 in the x-direction,
perpendicular to E (which is in the -y direction), and as they leave the region, they have an
angular deflection θ (see diagram). Neglecting end effects, show that
region, they have an angular deflection θ (see diagram).

+
θ
v0 d
y
L − x
Neglecting end effects, show that
e
 
VL
e
tan θ = . (1)
dv02 m
v0 Vacc

2eV
2. It appears from the equation derived in Question 1 above that the charge to mass ratio of the
electron can be determined in this way. However, in Thomson’s “cathode ray” apparatus, the
electrons were accelerated to their speed v0 across a potential difference Vacc between the
cathode and anode of the cathode ray tube. Show (trivially) that
r
2eVacc
ν0 = (2)
m
and hence that
V L
tan θ = (3)
Vacc 2d
which is independent of e /m!

3. Thomson overcame the difficulty by applying a magnetic field to the region between the plates.

(a) For a magnetic field perpendicular to the plane of the paper, show that the condition for
the electron beam to be undeflected is

e E2
= (4)
m 2Vacc B 2

(b) Should the magnetic field be directed into or out of the paper?
(c) Deduce from Eq.(4) that the ratio E /B has the units of velocity.
(d) If Vacc = 300 V, and the measured ratio of E /B is 1.027 × 107 m/s, deduce a value for
e /m.
(e) At what fraction of the velocity of light are the electrons travelling in this case?
(f) If V = 50 kV and d = 1 cm, what is the value of B?
(g) If E and B are held steady at these values, what happens (qualitatively) to the track of
the electron beam if the value of Vacc is increased above 300 V?

4. In the lectures, it was shown that a particle of charge q, moving at speed v in a uniform
magnetic field B perpendicular to the plane of the particle’s motion, follows a circular path of
radius
mv
r= (5)
qB
which it negotiates at a frequency (number of orbits per second)

qB
f= (6)
2πm
This is the principle of the device known as the cyclotron. Note that f , which is known as the
cyclotron frequency, is independent of the speed of the particle (at least under non-relativistic
conditions), and a function only of B and q/m!

(a) Find the radius of the track of a 2 MeV proton in a magnetic field of 0.2T.
(b) Find the cyclotron frequency for electrons in the same magnetic field.

Neglect relativistic effects.

5. Assessed problem A particle with charge q and mass m is free to move in a uniform magnetic
field B and a uniform electric field E. The magnetic field B is in the z direction and E is in the y
direction. Show that the equations of motion for the particle having a velocity v = vx î + vy ĵ + vz k̂
are
vx = vE + v⊥ sin(ωt + ψ) vy = v⊥ cos(ωt + ψ) vz = vk
for a suitable choice of v⊥ , vk , vE , ω and ψ. [3 marks]
The interaction of a very powerful laser beam with a solid target produces strong magnetic
and electric fields. Suppose the magnetic field is 200 Tesla (in the z direction) and the electric
field is 1010 Vm−1 (in the y direction). An electron is at rest at t = 0. What are the values of
v⊥ , vk , vE , ω and ψ for the motion of the electron? [2 marks]
Ignore relativistic corrections to the mass of the electron. Sketch vx and vy versus t. Describe
the motion of the electron [2 marks]

Answers to Exercises
A 9.4 × 10−5 m/s; 2 -

B 130J; 3 -; −z; -; 1.758 × 1011 C/kg; 3.4%; 0.487T;


downwards
C 0.15µV 4 r = 1.02m; f = 5.60GHz

1 - 5 -

Peter Török
Solutions  –  Problem  sheet  –  week  5  –  1st  year  E&M  

I 1
A. I = Nev A so v = = = 9.4 × 10 −5 m/s.
NeA 8.5 × 10 1.6 × 10 −19 π (5 × 10 − 4 ) 2
28

100 1.7 × 10 −8 60
B. Resistance R = Lρ / A ; Power = I 2 R ; Energy in 1 min = = 130 J.
π (5 × 10 − 4 ) 2
y
C. Bx
Δy x
Δx
The Hall voltage is registered across the y-dimension.

IBΔy IB
VHall = = since A = ΔxΔy
NeA NeΔx
3
VHall = = 0.15 µV.
8.5 × 10 1.6 × 10 −19 1.5 × 10 −3
28

1. Upward force on electron Fy = eE = eV / d


Upward acceleration a y = eV / md
eV L
Final upward speed v y = a yt =
md v0
where t = L / v0 is the time spent between the plates.
vy vy eVL
tan θ = ≡ = (1)
vx v0 mdv 02

2eVacc
2. 1
2
mv02 = eVacc so v0 = (2)
m
eVL m LV
tan θ = = (3)
md 2eVacc 2dVacc

3. (a) If the electric and magnetic forces are to balance


2
⎛E⎞ 2eV acc e E2
eE = ev 0 B ⇒ ⎜ ⎟ = v02 = ⇒ = (4)
⎝B⎠ m m 2Vacc B 2
(b) The magnetic forces needs to be in the downward (−y) direction, so v × B needs
to be in the +y direction, so B needs to be in the −z direction.
(c) This is obvious from eq.(4).
e (1.027 × 10 7 ) 2
(d) = = 1.758 × 1011 C/kg
m 2 × 300
The ratio e / m is now known to about 9 significant figures. To 6 figures, the value is
1.75882×1011C/kg so the value obtained in this question differs from the exact value
rounded to 4 figures by 1 in the last place.
v0 1.027 × 10 7
(e) = = 0.034 (3.4%)
c 3 × 10 8
E V /d 5 × 10 6
(f) = = 1.027 × 10 7 m/s, so B = = 0.487 T
B B 1.027 × 10 7

(g) If Vacc is increased, the magnetic force rises in relation to the electric force. So
the beam is deflected downwards (in the −y direction).

4. For the proton, 1


2 m p v 2 = 2× 10 6 e so v = 4 × 10 6 × (e / m p )
mpv m p 2 × 10 3
r= = = 1.02 m
eB e 0 .2
e 0.2
For the electron, f = =5.60 GHz.
m 2π
1st year E & M Biot-Savart law, Force on current element Problem sheet - Week 6

Useful constants and information:

• ε0 = 8.854 × 10−12 F m−1 ; • me = 9.109 × 10−31 kg;

• Atomic number density of Cu = 8.5 ×


• µ0 = 4π × 10−7 H m−1 ;
1028 m−3 .

• e = 1.602 × 10−19 C; • The resistivity of copper is 1.7 × 10−8 Ωm.

Some simple exercises to get you going


(Answers are given at the end)

A. Find the force per unit length between two long straight parallel wires, 5 cm apart, carrying
currents of 1 kA in opposite directions. Is the force attractive or repulsive?

B. What is the magnitude of the magnetic field experienced by each wire due to the other in Part
A above?

C. Find the current in a 30 turn toroidal solenoid of radius 8 cm needed to produce a magnetic
field of B = 3.0 × 10−3 T at its centre.

D. A power line running east-west 20 m above the ground carries a current in an easterly direc-
tion. What is the direction of the associated magnetic field? Find the current in the wire for
which the magnitude of the magnetic field at ground level directly below the line is roughly the
same as that of the earth’s field (∼10−4 T).

Problems
1. A straight piece of wire running along the y-axis between y = ±L /2 forms part of an electrical
circuit carrying current I (see Fig. 1 below). Show that its contribution to the magnetic field at
P, a distance a from the origin in the x − z plane, is

µ0 I cos θ
B=
2πa
HINT: Use a simple extension to the proof for the infinite straight wire given in the lectures.
There, we integrated over an angle θ, but you can equally integrate over y using the standard
integral: Z
dy 1 y
= .
(a 2 + y 2 )3/2 a 2 (a 2 + y 2 )
p
y

x
I
a P X
L
r

θ
θ
2b b

Fig. 1 Fig. 2

2. A current of 3 A flows anticlockwise round the trapezoidal circuit shown in fig. 2. If b = 4 mm


and θ = 45◦ , find the magnitude
= 45! and direction of the magnetic field at X.

3. A square current loop of side a centred at the origin has its four corners at x, y = ±a /2. The
loop carries a current I in a clockwise sense for an observer looking in>>the
t +z direction. Show
that the magnetic field on axis at (0, 0, z) where z  a is
 µ0 m 
B = 0, 0,
t ak 2πz 3

where m (= Ia 2 ) is the magnetic dipole moment of the loop. Compare the answer with the
analogous result for an electric dipole.
[HINT: Use the Biot-Savart Law to find the field from a single side. Don’t forget to resolve
along the z-axis, and to use the fact that z  a. Since a is small, you don’t need to integrate.
Multiply by 4 to get the result from all four sides.]

4. ***

(a) Show that a particle of mass M and charge q moving in a circular orbit with (vector)
angular momentum L is equivalent to a magnetic dipole with a dipole moment m =
qL/2M.
(b) In the Bohr model of the hydrogen atom, the orbital angular momentum of the electron
is restricted to integer multiples of ~ (= h /2π). Calculate the magnitude of the magnetic
dipole moment of the electron in the ground state of the hydrogen atom. [This value is
termed the Bohr magneton, and is usually denoted µB .]
(c) The potential energy of a magnetic dipole m in a magnetic field is U = −m · B. If a
magnetic field is applied to a collection of hydrogen atoms, one might expect the atoms to
line up in the direction of lowest potential energy (i.e. with their dipole moments aligned
in the direction of B). Calculate the potential energy difference between orientations
parallel and antiparallel to the magnetic field for a single hydrogen atom in a magnetic
field of 10 T (which is about the maximum steady magnetic field that can be generated
in the laboratory). Compare this with the average thermal kinetic energy of hydrogen
atoms (∼ 32 kT at room temperature). What conclusion can be drawn?

5. Assessed problem

(a) The Law of Biot and Savart gives the differential contribution dB to the magnetic field at
a point P from a current element Idl as
µ0 Idl × r̂
dB =
4π r 2

where r = r r̂ is the position of P relative to the current element. Use this law to show
that the magnitude of the magnetic field on the axis of a circular loop of radius a at a
distance h from the centre is
µo Ia 2
B= 3/2
2 a2 + h2
where I is the current flowing in the loop. [10 marks]
(b) The centres of a pair of Helmholtz coils, each of radius a and carrying a current I (in the
same direction), lie on the x-axis at x = ±a /2 as shown in the following diagram:

Determine the magnitude of the magnetic field at a general point on the axis. [5 marks]
(c) i. Find the value of the magnetic field at the mid-point (x = 0) between the two
Helmholtz coils.
ii. Determine the value of dB
dx at x = 0
iii. Determine the value of the magnetic field at x = ±a /2
iv. Plot the magnetic field along the x axis for −a /2 ≤ x ≤ a /2.
[10 marks]

Answers to Exercises
A 4.0 N/m repulsive; 2 70.0µT
B 4 × 10−3 T 3 -
C 40 A 4 (a) -; (b) m = 9.28 × 10−24 Am2 ;
D North. 10 kA (c) ∆U = 1.86 × 10−22 J

1 - 5 -

Peter Török
Solutions  –  Problem  sheet  –  week  6  –  1st  year  E&M  

µ 0 I 2 4π × 10 −7 10 6
A. f = = = 4 N/m.
2π d 2π 0.05
µ I
B. B = 0 = 4 × 10 −3 T.
2πd
µ IN 2! aB 2 ! ! ! 0.08 ! 3 !10 "3
C. B= 0 so I = = = 40A
2! a µ0 N 4! !10 "7 ! 30
2πrB 2π 20 × 10 −4
D. B points north. I = = =10 kA.
µ0 4π × 10 −7

+L/2
µ Ia + L / 2 dy µ Ia 1 ⎡ y ⎤ µI L µ I cosθ
1. B= 0 ∫
4π − L / 2 (a + y )
2 2 3/ 2
= 0
4π a 2
⎢ ⎥ = 0
⎢⎣ (a 2 + y 2 ) ⎥⎦ − L / 2 4π a (a 2 + L2 / 4)
= 0
2π a

µ0 I
→ as L → ∞ and θ → 0 ,
2π a
µ 0 I cos 45o ⎛ 1 1 ⎞ 4π ×10−7 × 3 2
2. B= ⎜ − ⎟= = 70.7 µT directed into the paper.
2π ⎝ b 3b ⎠ 2π 2 4 ×10−3 3
3.
I B1

r θ
a
z

µ0 Ia
Magnetic field from top side of square B1 ≅ where z ≅ r for z ! a . Resolving
4π z 2
this field along the z-direction and multiplying by 4 (because the square has four sides)
µ0 Ia a µ0 Ia 2 µ0 m
yields Bz = 4 = = .
4π z 2 2 z 2π z 3 2π z 3

qv qvπa 2 qL
4. (i) m = IA = Iπa . But I =
2
and L = Mva . So m = = .
2πa 2πa 2m
e 1.6 ×10 −19 6.63 ×10 −34
(ii) m= = = 9.28 ×10 −24 A m2.
2me 4π 9.1×10 −31

(iii) ΔU = 2mB = 1.86 × 10 −22 J.


1st year E & M Ampère’s law and induction Problem sheet - Week 7

Useful constants and information:

• ε0 = 8.854 × 10−12 F/m; • me = 9.109 × 10−31 kg;

• e = 1.602 × 10−19 C; • µ0 = 4π × 10−7 H/m

Some simple exercises to get you going


(Answers are given at the end)

A. A solenoid 1 cm in diameter and 10 cm long has 1000 turns. Find the magnetic field in the
solenoid if it carries a current of 500 mA. (Assume the “long solenoid” formula applies.)

B. Use the principle of superposition to show that the magnetic field at the end of a long solenoid
with n turns per unit length and carrying a current I is µ0 nI/2.

C. A loop of wire has area 5cm2 (= 5 × 10−4 m2 ) and resistance 3kΩ. If a uniform magnetic field
of 0.6 T normal to the plane of the loop is switched on, how much charge flows in the circuit?

Problems
1. In a basic dynamo, a rectangular current loop of area A connected to a circuit of resistance R
is rotated at angular velocity ω about a uniform magnetic field B. The magnetic flux through
the loop is ΦB = BA cos ωt.
Find expressions for the EMF generated, and the current in the circuit. Show that the time-
averaged power dissipated in the resistor is (BA ω)2 /2R.
Show that the couple opposing motion is Γ = IAB sin ωt, and hence verify that the power
needed to rotate the loop is the same as the power dissipated in the resistor.

2. A coaxial cable consists of two concentric cylindrical conductors of radii a and b(> a) carrying
current I in opposite directions. Use Ampère’s Law to show that the magnetic field between
the conductors is given by
µ0 I
B=
2πr
3. A square current loop of side a centred at the origin has its four corners at the origin has its
four corners at x, y = ±a /2. The loop carries a current I in a clockwise sense for an observed
looking in the +z direction. Show that the magnetic field at (x, 0, 0) where x  a is
 µ0 m 
B = 0, 0, −
4πx 3

where m(= Ia 2 ) is the magnetic moment of the loop.

4. A thick current sheet occupies the region −a < z < a and is infinite in extent in the x and
y-directions. It carries a uniform current density j = j0 î. Calculate the magnetic field outside
the sheet for z > a. Also calculate the field for 0 < z < a, i.e. inside the sheet.
5. An inductor is formed from a cylindrical air-core solenoid of length `, radius a, and N turns of
wire. Calculate the inductance of the solenoid.

6. Assessed problem A square loop of wire, side a, mass m, resistance R, and lying in the
plane of the paper, falls under gravity through a region of depth w (> a) in which there is a
uniform magnetic field of strength B directed into the paper (see diagram). The loop starts
from rest at t = 0 with its leading edge at the boundary of the field region (as shown).

(a) Demonstrate that the speed of fall v is initially governed by the differential equation
dv (aB)2
= g − bv where b =
dt mR
(b) Hence show that
g
v= (1 − exp(−bt))
b
(c) How does the situation change after the trailing edge of the loop enters the region of the
field?

[15 marks]

Answers to Exercises
A 6.3mT; 3 -
B -
4 −µ0 j0 a ĵ, −µ0 j0 z ĵ
C 100nC
5 N 2 πa 2 µ0 /`
1 -
2 - 6 -

Peter Török
st
Solutions – Problem sheet – week 7 – 1 year E&M

A.
B = µ0 nI = 4π 10 −7 10 4 0.5 = 6.3 mT.

B.
Consider a long solenoid in two halves P

1 2

By symmetry, the contributions of 1 and 2 at P are equal. So each on its own


delivers half the total.

dq dΦ B 0.6 5 ×10 −4
∫ dq = R ∫
−1
C. emf = IR = R = . Hence dΦ B = BA / R = = 100
dt dt 3000
nC.

dΦ B BAω sin ωt ( BAω ) 2 sin 2 ωt


1. V =− = BAω sin ωt ; I = ; W ≡ VI = ;
dt R R
W = ( BAω ) 2 / 2R

using the fact that the time-average of sin2 is 0.5.

Assume that the loop rotates about a vertical axis, the vertical sides of the loop are
of length b, and the horizontal dimension of the loop is of length 2a. The force on
each vertical side is IbB and the moment of the force about the axis is IbBa sin ωt .
Since there are two vertical sides, and A = 2ab , it follows that the torque is
Γ = IBA sin ωt .

The work that must be done against this torque to rotate the loop is
dθ BAω sin ωt
W =Γ = Γω = × BAω sin ωt leading to the same result as before.
dt R

2.

r
a
b

µ0 I
Field at radius r =
2πr
b
µ0 Idr µ0 I
For a cable of length  , Φ B =

a
2πr
=

log{b / a}

L Φ B µ0
Self inductance per unit length≡ = log{b / a}
 I 2π
2
B2 1 ⎛ µ0 I ⎞
Energy per unit volume = = ⎜ ⎟
2µ 0 2µ 0 ⎝ 2πr ⎠
b 2
⎛ µ0 I ⎞
1 µ I2
Energy per unit length =
a

2µ 0 ⎝ 2πr ⎠ ∫
⎟ 2πrdr = 0 log{b / a}.

3. Looking along the +z-axis ....... (x points left to ensure a right-handed system)
A
y B  

x
D
C  
VERTICAL SIDES (AD and BC)

Contribution to B at P is OUT OF the paper (−z direction)


µ0 ⎧ Ia Ia ⎫ µ 0 Ia ⎛ a a ⎞ µ 0 Ia 2
Bxv = − ⎨ − ⎬ ≅ − ⎜1 + − (1 − ) ⎟ = −
4π ⎩ ( y − a / 2) 2 ( y + a / 2) 2 ⎭ 4πy 2 ⎜⎝ y y ⎟⎠ 2πy 3

HORIZONTAL SIDES (AB and CD)

Contribution to B is INTO THE paper (+z direction)


µ 0 Ia a / 2 µ 0 Ia 2
B =2
h
x =
4πy 2 y 4πy 3
Adding the vertical and horizontal contributions yields
µ 0 Ia 2 µ 0 Ia 2 µ 0 Ia 2 µ m
Bztotal = − + = − = − 0 3 where m is the magnetic moment.
2πy 3
4πy 3
4πy 3
2πy

4.
5

Das könnte Ihnen auch gefallen