You are on page 1of 44

Chapter 15

Electrostatics
Electrical forces 15.1 Forces between charged particles
predominate in the Positive and negative charges
interaction between the Coulomb's law

atoms and molecules 15.2 The electric ®eld


of ordinary matter. This Addition of electric forces
chapter explains the De®nition of the electric ®eld
Lines of force
concepts of electric
Superposition of electrostatic ®elds
®eld and electrostatic
15.3 Gauss's law
potential that are
Flux
needed to understand
Surface integrals
the behaviour of these
15.4 The electrostatic potential
forces.
Work done by electric charges
Equipotential surfaces
The dipole potential and ®eld

15.5 Electric ®elds in matter


Macroscopic electric ®elds
Conductors in electric ®elds
Insulators in electric ®elds
Polar molecules

15.6 Capacitors
Relative permittivity
Stored energy
Energy density of the electric ®eld
542 1 5 : ELE C TRO S TA TI CS

Everyone is familiar with electricity as a source of power. Pressing a


switch will turn on a light or heat an oven. Energy is continuously being
produced in these processes, energy that is carried by an electric current
through the metal wires connected to the electricity supply. The electric
current is made up of a ¯ow of moving electrons. We cannot see the
movement because the electrons are very small and are able to move
through a metal without disturbing the structure of the metal.
Electrons are pushed along a wire by forces that act on them because
they carry electric charge. These forces are called electric forces. The
electric force on a charged particle is the same whether the charge is
stationary or moving. There are additional forces that act only on moving
charges, which are called magnetic forces. Moving charges and magnetic
forces are discussed in Chapter 16. The subject of this chapter is
electrostatics, which is the study of the electric forces acting on stationary
charges, and of how these forces are modi®ed in the presence of matter.
Like gravitational forces, electrostatic forces act at a distanceÐthere is
an electrostatic force between two charged particles even if they are
separated by a vacuum. The magnitude of the electrostatic force also has
the same inverse square variation with distance as the gravitational force.
There are, however, two very important differences between gravitational
and electrostatic forces. The ®rst is that the gravitational force between
two masses is always attractive, whereas charges may attract or repel one
another. The other difference is that, on an atomic scale, electrostatic
forces are enormously strong compared with gravitational forces. In the
discussion of the internal motion of individual atoms and molecules,
which is the subject of Chapter 11, only electrostatic forces were
considered and the effects of gravitational forces were completely
neglected. This seems paradoxical, because in everyday life we are well
aware of gravitational force, but do not often notice electrical forces. The
reason is that electrostatic attractions and repulsions tend to cancel out,
whereas gravitational forces are always attractive and, in particular, the
whole of the Earth attracts everything on its surface.
Although the laws governing the forces between charges are introduced
in this chapter in the context of electrostatics, these laws always apply,
even when magnetic effects or electromagnetic waves are present. The
chapter starts by discussing the forces between very small idealized electric
charges in order to explain the concepts of electric ®eld and potential.
Later on, in Sections 15.5 and 15.6, we are concerned with objects
containing very large numbers of atoms. Only average electrical
properties are then of interest: it will be shown how these averages can
be obtained without having to consider the electrical forces within each
atom in turn.
FORC E S BE T WE EN CH ARG ED P AR T IC LE S 543

15.1 Forces between charged particles

Positive and negative charges


It was mentioned above that electrostatic forces are sometimes attractive
and sometimes repulsive. This is because there are two different kinds of
charge, which are called positive and negative. Just like positive and
negative numbers, positive and negative charges are described as being of
opposite sign. Electrons carry a negative charge. Like numbers, positive
and negative charges may cancel one another out. For example, as is
described more fully in Chapter 9, an atom consists of a number of
electrons bound to a positively charged nucleus. The charge on the
nucleus has exactly the same magnitude as the charge of all the electrons
in the atom. Since the nuclear charge is of opposite sign to the charges on
the electrons, the net charge carried by the atom, which is the algebraic
sum of all its charges, is zero. The atom is said to be electrically neutral.
In SI units, charge is measured in coulombs (symbol C). The coulomb
is de®ned with reference to the force between wires carrying electric
current: this is discussed in Chapter 16. The charge carried by a single
electron is written as ÿe, and its magnitude is
e ˆ 1:602  10ÿ19 coulombs
to four signi®cant ®gures. Because the coulomb is a very large unit,
charges are often measured in microcoulombs (symbol mC:
1mC  10ÿ6 C).
Ordinary matter, made up of electrons and nuclei, may be electrically z All observable charges are
neutral or may have a charge that is e times an integer. Other particles multiples of the electronic
besides electrons and atomic nuclei are found in cosmic rays, or may be charge
created in high-energy collisions in accelerators. All these particles also
have charges that are zero or e times an integer. Within a nucleus there
are thought to be particles called quarks that carry an amount of charge
that is a fraction of e. However, quarks have a property called
con®nement, which means that they are never observed singly but go
around in packets that do not have fractional charge. It is thus a universal
rule that any object is either electrically neutral or has a positive or
negative charge with magnitude that is an integral multiple of e.

Coulomb's law
Coulomb's law tells us the strength and direction of the forces acting
between two charges. This is the simplest possible case, but on the basis of
Coulomb's law it is possible to work out the electric force on any
distribution of charges. Consider two charges that we shall label q1 and q2 :
q1 and q2 are numbers representing the magnitudes of the charges in
544 1 5 : ELE C TRO S TA TI CS

coulombs, and these numbers are positive or negative depending on


whether the charges are positive or negative. We shall idealize the
problem by supposing that q1 and q2 occupy such a small volume that
they may be treated as point charges with no spatial extent at all.
Let us de®ne F12 to be the magnitude of the force exerted by a charge q1
on a charge q2 . According to Coulomb's law, F12 depends on the product
q1 q2 ; doubling the magnitude of either charge doubles the strength of the
force. How do the forces between q1 and q2 vary with distance? Just like
the gravitational force between two masses, the electrostatic force between
two charges varies as the inverse square of their distance apart. This
inverse square law is known to hold with great precision, not from direct
measurement of the force between charges, but from the observation of
other phenomena that are deduced from the inverse square law. The
magnitude of the force between the charges q1 and q2 is expressed
mathematically as
q1 q2
F12 / 2 : …15:1†
r12
A constant of proportionality is required to give the strength of the force
in newtons when q1 and q2 are in coulombs and r12 in metres. In the
SI system the constant is written as 1=4p0 Ðincluding the factor 4p
simpli®es other equations in electricity and magnetism. The equation for
the magnitude F12 of the force becomes
q1 q2
F12 ˆ 2 : …15:2†
4p0 r12

The direction of the force between the two charges depends on their
signs. The force acts on the line joining the charges and, if the charges
have opposite sign, like the negatively charged electron and the positively
charged nucleus in the hydrogen atom, the force is attractive: it is the
electrical attraction that binds an electron to the nucleus and ensures that
the atom is stable. On the other hand, if both charges are positive, or both
are negative, the force between them is repulsive.
We must be careful to get the direction of the force correct when
setting up the ®nal equation for Coulomb's law. To do this we must use
the vector notation, and in particular we shall use unit vectors, which are
vectors pointing in any direction, but which always have unit length.
Since we are interested in the direction between the two charges, we
introduce the vector ^r12 , which is a vector of unit length pointing from an
origin at the centre of charge q1 towards charge q2 .

Unit vectors are used in Section 20.2 (in the mathematical review at the
end of the book) to de®ne the directions of the axes of a Cartesian
coordinate system. In that context, the unit vectors i, j, and k (all of unit
FORC E S BE T WE EN CH ARG ED P AR T IC LE S 545

length) are pointing in the ®xed directions chosen for the x-, y-, and z-
axes. Any vector a that has components ax , ay , and az along the x-, y-,
and z-axes can then be written as ax i ‡ ay j ‡ az k. This expression
speci®es both the magnitude and direction of the vector a. Here it is more
convenient to use a different notation, allowing unit vectors to point in
any direction. The symbol ^is used to indicate that a vector has unit length:
thus ^a is a vector of unit length pointing in the same direction as a.

The force exerted by charge q1 on charge q2 is denoted by the vector


F12 . This force points along the line joining the charges, in the direction
away from q1 for a repulsive force (q1 and q2 having the same sign) and
towards q1 for an attractive force (q1 and q2 having different signs).
Different possibilities are illustrated in Fig 15.1, which shows both F12 and
^r12 for different signs of the charges. In Fig 15.1(c), where the signs are
different, the force is towards q1 , which is in the opposite direction to ^r12 .
However, the unit vector ÿ^r12 is also in the opposite direction to ^r12 . The
sign required in front of the unit vector ^r12 is thus the same as the sign of
the product q1 q2 . Fig. 15.1 The force on a charge due
The force between two charges q1 and q2 may now be expressed in to the presence of another charge is
mathematical terms, using the same notation as in Fig 15.1 for the in the same direction as the unit
position vector of q2 with respect to q1 . The magnitude of the force is vector on the line joining the
given by eqn (15.2) and in SI units Coulomb's law is charges. (a), (b), and (c) show the
q1 q2 force on charge q2 caused by q1 for
F12 ˆ ^
2 r12 : …15:3† different combinations of the sign of
40 r12
the charges. (d) shows the force on
Similarly, the force F21 exerted by q2 on q1 is q1 caused by q2 when both are
q1 q2 positive.
F21 ˆ ^
2 r21 : …15:4†
4p0 r12
Since ^r21 is a unit vector pointing from q2 towards q1 , in the opposite
direction to ^r12 , the forces exerted on the two charges according to
eqns (15.3) and (15.4) are equal and opposite, as they should be (compare
Figs 15.1(a) and (d)).
In words, Coulomb's law states that
The force between two charges acts along the line between the
charges, and is proportional to the product of the charges and to the
inverse square of the distance between them. The force is repulsive for
charges of the same sign and attractive for charges of opposite sign.
The dimensions of all the quantities in eqn (15.3) are de®ned
independently of Coulomb's law. The unit of force, the newton, is de®ned
by Newton's second law. The unit of charge, the coulomb, is de®ned with
reference to the magnetic force between wires carrying current. To satisfy
eqn (15.3), the units of the constant 0 are C2 Nÿ1 mÿ2 . Its value is related
546 1 5 : ELE C TRO S TA TI CS

to the speed of light, which is the distance light travels in a vacuum in one
second. Since the the unit of length is itself de®ned in terms of the time
taken for light to travel a distance of one metre, the speed of light is also
de®ned to be a particular number of metres per second. Scientists all over
the world have agreed that the value of the speed of light is exactly
2:997 924 58 m sÿ1 . Because the constant 0 , which is called the
permittivity of free space, is derived from the speed of light, it is also
in principle known exactly. However, it is not a rational number, but
when expressed in decimals it can be calculated to any number of places.
To four signi®cant ®gures, its value is
0 ˆ 8:854  10ÿ12 C2 Nÿ1 mÿ2 : …15:5†

Worked Example 15.1 Two small particles of carbon, each weighing 1 mg


and each carrying a charge of 10ÿ6 C, are one centimetre apart. Calculate
the electrostatic force between them.
Answer The force between the particles is found directly by substitution
in eqn (15.2). It is
10ÿ6  10ÿ6 10ÿ8
ÿ4
  90 N:
4p0  10 1:1  10ÿ10
This example illustrates the enormous strength of electrostatic forces.
The force of 90 N is nearly equal to the weight of a 10 kg mass, acting
between two tiny particles. If the particles were free to move, their initial
acceleration would be 90 km sÿ2 . In practice it is not possible to
accumulate as much charge as 1 mC on such small pieces of matter, even
though only a small fraction of the atoms need to gain or lose an electron
to reach this value. The number of atoms in one mole of carbon is
Avogadro's number, NA  6  1023 , and, since the mass number of
carbon is 12, the number of atoms in 1 mg is about 5  1019 . The number
of electronic charges in 1 mC is 10ÿ6 =e  1013 =1:6. The fraction of carbon
atoms that must lose one electron to charge the particles with 1 mC is
1013 =…5  1019  1:6†, or a little more than one in a million.

15.2 The electric ®eld

Coulomb's law in the form given in eqn (15.3) enables us to work out the
forces that two point charges exert on each other. Most practical electrical
problems involve not just two charged particles, but vast numbers of
them. This section introduces the idea of the electric ®eld, which describes
the force on a charged particle due to all the other charges in its
neighbourhood.
T HE ELE C TRI C FIE LD 547

Addition of electric forces


In Fig 15.2 a third positive charge q3 has been brought close to the two
positive charges q1 and q2 shown in Fig 15.1(d). There is now an
additional force F31 acting on q1 , due to q3 . However, the force F21
exerted on q1 by q2 is unchanged, provided that q1 and q2 remain at the
same positions after q3 has been introduced. This is described by saying
that the electric force between charges is a two-body force: the force
between two charges is given by Coulomb's law independently of the
presence of any other charges. The same applies to q1 and q3 , and the
force F31 is also given by Coulomb's law. The net force F1 on q1 is simply
Fig. 15.2 The net force F1 on the
the vector sum of the forces due to q3 and q2 separately,
charge q1 is the vector sum of the
forces F21 and F31 caused by q2 and
F1 ˆ F21 ‡ F31 : …15:6†
q3 . In the diagram F1 is the diagonal
in a parallel of forces.
A very simple application of eqn (15.6) is to two or more charges that
are very close together. For example, suppose that in Fig 15.2 q2 and q3
both are of magnitude ‡e, and q3 is moved to the same location as q2 .
2
The force on q1 then has a magnitude 2eq1 =4p0 r21 and is in the direction
^r21 . This is, of course, the same as the force due to a single charge 2e at the
position of q2 . The fact that the force is a two-body force is thus already
included in Coulomb's law, which allows q1 and q2 to have any values,
although we know that in reality the charge in a small volume is always
built up of individual charges of magnitude e.
As more and more charges are added, each exerts a force on all the z The force on a charge is the
others. Labelling the charges one by one as q1 , q2 , q3 , . . . qj , . . . , the force vector sum of the forces due to
Fi on a particular charge qi is the vector sum of the forces Fji due to all the all other charges
others,
X X qi qj qi X qj
Fi ˆ Fji ˆ 2
^rji ˆ ^rji : …15:7†
j6ˆi j6ˆi
4p0 rji 4p0 j6ˆi rji2

The caption j 6ˆ i under the summation signs indicates that the sum is
taken over all values of j except j ˆ i, since the charge qi is not exerting a
force on itself. All the unit vectors ^rji in the equation remind us that the
force between each pair of charges is pointing along the line joining the
charges. However, when doing calculations it is usually convenient to
refer all the position vectors to a ®xed origin rather than dealing with
each pair of charges separately. Figure 15.3 shows two charges qi and qj
with position vectors ri and rj referred to an origin at O. The vector from
qj to qi is rji ˆ ri ÿ rj . Writing the length of this vector as jri ÿ rj j, the
unit vector in the direction from qj to qi is
Fig. 15.3 The vector rij between
ri ÿ rj the charges qi and qj is the differ-
^rji ˆ : …15:8† ence of the position vectors ri and rj .
jri ÿ rj j
548 1 5 : ELE C TRO S TA TI CS

Substituting in eqn (15.7), the force on qi becomes

qi X qj …ri ÿ rj †
Fi ˆ : …15:9†
40 j6ˆi jri ÿ rj j3

Note that, because each term in this expression has a vector of length
jri ÿ rj j in the numerator, the length appears raised to the power 3 in the
denominator, even though the force follows an inverse square law.

Worked Example 15.2 Four positive charges, each of magnitude q, are


situated at the corners of a square of side a, as shown in Fig 15.4. What is
the magnitude and direction of the force on the charge at A?
Answer Consider the components of the force in the directions BD and
CA. The charges at B and D give rise to equal and opposite components
along the direction BD, and each has a component
q2 cos…45 † q2
2
ˆ p
4p0 a 4 2p0 a2

Fig. 15.4 The charges at A, B, C,


along the direction CA. The force due to the charge at C is along CA and
and D lie in the same plane at the has a magnitude
corners of a square of side a: q2 q2
p 2 ˆ
4p0 … 2a† 8p0 a2 :
p
The total force on the charge at A is therefore q2 …1 ‡ 2 2†=…8p0 a2 †
pointing along the direction CA. Each of the other charges has a force of
the same magnitude pointing in the direction of the outward diagonal,
and the net force on the square is zero.

Exercise 15.1 Calculate the magnitude of the force on the charge at A in


Fig 15.4 if the charge at B is replaced by a charge ÿq.
Answer The force due to the charges ‡q at D and ÿq at B is now
p
2 2q2 =…8p0 a2 † in the direction DB, and the total force has magnitude
3q2 =…8p0 a2 †.

De®nition of the electric ®eld


Imagine that you have a test charge q that you can move about in the
region near the charges qi . Suppose also that the positions of the charges
qi are undisturbed by the presence of q. If q is placed at a point with
position vector r with respect to the origin at O, the force on it is,
T HE ELE C TRI C FIE LD 549

according to eqn (15.9),


q X qj …r ÿ rj †
Fˆ : …15:10†
4p0 j jr ÿ rj j3

The sum is over all j now, because the test charge has been excluded
from the labelling. Now look at the quantity F=q. It does not depend on
the test charge at all. It may be calculated at any location whether or not a
test charge is present; the position vector r is a variable, and F=q is a
vector function of position. This function is called the electric ®eld, and
it is denoted E…r†.

Functions of position are called ®elds. Because E…r† is itself a vector, it is a


vector ®eld. In electromagnetism we shall also meet functions of position z Vector ®elds
that are scalar quantities, which have magnitude but not direction. These
functions are called scalar ®elds.

By dividing eqn (15.10) by q we ®nd


1 X qj …r ÿ rj †
E…r† ˆ : …15:11†
40 j jr ÿ rj j3

The dimensions of the electric ®eld are force per unit charge, and it is
measured in newtons per coulomb: if a charge of one coulomb were
placed in an electric ®eld of strength one newton per coulomb it would
experience a force of one newton, and this force would act in the
direction of the electric ®eld vector at the position of the charge.

Worked Example 15.3 Calculate the electric ®eld due to a proton at a


distance of 0.07 nm (this distance is approximately the separation of the
protons in a hydrogen molecule).
Answer There is only a single term in the summation in eqn (15.11) and
we can choose the origin to coincide with the proton. At any point a
distance r ˆ 0:07 nm from the proton, the electric ®eld points in a
direction away from the proton, and has a magnitude e=…4p0 r 2 †:
1:602  10ÿ19
4p  8:854  10ÿ12  0:0049  10ÿ18
 2:9  1011 newtons per coulomb:
This is far in excess of any electric ®eld that can be achieved over a
distance larger than atomic dimensions, and again illustrates the
enormous strength of electric forces within atoms and molecules.
550 1 5 : ELE C TRO S TA TI CS

Exercise 15.2 Estimate the repulsive electrostatic force between the two
protons in a hydrogen molecule and compare it with their gravitational
attraction.
Answer The electrostatic force is about 5  10ÿ8 N, and the gravitational
force is about 4  10ÿ44 N. The electrostatic force is thus more than 1036
times larger: this ratio applies at any distance, since both forces obey an
inverse square law.

Lines of force
Around an isolated positive point charge q1 there is only a single term in
the summation in eqn (15.11). Choosing the origin to be at the position
of the point charge, the electric ®eld is
q1 r
E…r† ˆ : …15:12†
4p0 r 3
In this equation r is the position vector of any point in space with respect
to the origin. The electric ®eld E…r† is everywhere pointing in the same
Fig. 15.5 Electric ®eld lines radiate direction as r, directly away from the origin. A positive charge q placed at
outwards from a positive point r will experience a force qq1 =…4p0 r 2 † in this direction. This can be
charge and inwards towards a visualized by drawing lines of force in the direction of the ®eld, as in
negative point charge. Fig 15.5. The diagram is only two-dimensional, but it will look the same in
any plane passing through q1 . The diagram does not indicate the strength
of the force, but notice that close to q1 , where the ®eld is strong, the lines
are close together, whereas the lines are far apart at large distances where
the ®eld is weak. The ®gure also shows the ®eld around a negative point
charge. The diagram is the same except that the ®eld is in the opposite
direction, inwards instead of outwards. Following the arrows, you can see
that ®eld lines start from positive charges and end on negative charges.
An instructive diagram of lines of force is shown in Fig 15.6. Here two
charges q of the same magnitude but opposite sign are placed not very
far apart. Close to each charge, the lines behave in the same way as in
Fig 15.5, pointing away from the positive charge and towards the negative
charge. But, as the distance from one charge increases, the in¯uence of the
other becomes more important. Field lines leaving the positive charge
bend round and move towards the negative charge. At larger distances
Fig. 15.6 The ®eld lines around an from the charges, the lines are far apart. The electric ®eld has become
electrostatic dipole start at the weak because the contributions from the positive and negative charges
positive charge and bend round to almost cancel one another. Once again the diagram is two-dimensional,
end on the negative charge. but it will look the same in any plane passing through both charges.
The pair of equal and opposite charges separated by a small distance is
z The electric dipole called an electric dipole. The ®eld pattern generated by an electric dipole
G A USS' S LAW 551

is important in many branches of physics and chemistry; later on we shall


evaluate the ®eld due to an electric dipole mathematically. However, it is
also very helpful towards understanding the behaviour of electric ®elds to
have a mental picture of the kind given by diagrams of lines of force. As in
the example of the electric dipole, these often illustrate important pro-
perties of the electric ®eld without the need to do any calculations at all.

Exercise 15.3 Sketch the lines of force in the neighbourhood of two


positive charges of equal magnitude.

Superposition of electrostatic ®elds


The electric ®eld as given by eqn (15.11) is the vector sum of the electric
®elds generated by each charge qj separately. If some more charges are
added, more terms are added to the summation. However, there is no
change to the terms that were already there, provided that the original
charges do not move. If we know the electric ®elds generated by two
different sets of charges separately, the electric ®eld generated by both
together is simply the vector sum of the two separate ®elds. The two
®elds, which each occupy three-dimensional space, are superimposed on
one another. Because it has this property, the electric ®eld is said to satisfy
the principle of superposition.
Superposition is discussed for one-dimensional waves in Section 6.5,
where it is shown that different waves may be superposed because the
equations governing the wave motion are linear. Similarly here, super-
position applies to different electric ®elds because the ®eld depends
linearly on the charges that generate the ®elds. When superposition is
extended to the varying electric and magnetic ®elds that occur in electro-
magnetic waves, the phenomena of diffraction and interference described
in Sections 8.6 and 8.7 can be explained.

15.3 Gauss's law

The electric ®eld due to any system of charges is found by superposing the
®elds due to each one separately. This sounds very simple but, since the
®elds to be summed are vectors, the general expression given by
eqn (15.11) may be very dif®cult to work out.
A completely different way of relating the electric ®eld to the charges is
called Gauss's law. It is sometimes much easier to calculate the ®eld from
Gauss's law than by summing the ®elds from all the charges. Gauss's law
552 1 5 : ELE C TRO S TA TI CS

follows from the inverse square variation of the electric ®eld with
distance, and it can be understood by analogy with the spreading out of
energy from a light source, which also decreases with the inverse square
of distance. If the light source is in an enclosed space such as a room, all
the light leaving the source reaches a surface somewhere in the room.
Nearby surfaces are brightly illuminated and those that are far away are
dimly illuminated. Moving the surfaces will make a difference to their
brightness, but not to the total amount of light in the room. We shall
prove that a similar result holds for a quantity called the ¯ux of the
electric ®eld. If a charge is surrounded by a closed surface, the ¯ux over
the whole surface has a ®xed relation to the amount of charge,
independent of the shape or size of the surface.

Flux
Figure 15.7 shows a small ¯at surface of area dS placed in an electric ®eld
E so that the normal to the surface is at an angle  to the direction of E.
Fig. 15.7 The vector dS has a
magnitude equal to the area dS
The projected area of dS viewed along the direction of the ®eld lines of
of the small surface and is E is dS cos . The ¯ux of E through dS is de®ned to be EdS cos . This
perpendicular to it. may be expressed concisely by associating a vector dS with the area dS,
directed along the normal to the surface, as shown in Fig 15.7. The ¯ux
through dS can now be written as the scalar product E  dS. Note that dS
can be the normal to the surface in either direction from the surface. The
sign of the ¯ux depends on whether  is greater or less than 90 . If the
component of dS in the direction of the ®eld is positive,  is less than 90
and the ¯ux is positive: if this component is opposite to the ®eld, the ¯ux
is negative.
If we have a large area S, which is not necessarily plane, we can divide it
up as shown in Fig 15.8. If the division is ®ne enough, the small surfaces
like dSi are practically ¯at. The ¯ux through dSi is E…ri †  dSi , where ri is
the position vector of dSi , the total ¯ux through S is the sum
P
i E…ri †  dSi of contributions from all the surfaces dSi . In the limit as
the areas of all the surfaces dSi tend to zero, the summation becomes a
two-dimensional surface integral over the surface S which is written as
X Z
Flux through S ˆ lim E…ri †  dSi ˆ E…r†  dS: …15:13†
dSi !0 S
i

Fig. 15.8 Any surface like the How does this equation apply to the electric ®eld around an isolated
shaded surface S may be divided up point charge q1 located at the origin of coordinates? Choose for the
into many adjacent surfaces dSi . In surface S a sphere of radius r centred at the origin. The electric ®eld on
the limit as the dSi become in®nite- the surface of the sphere has a magnitude q1 =…4p0 r 2 † and is
simal, each one may be regarded as perpendicular to the surface, so that the outward normal to the sphere
a plane surface. is everywhere in the same direction as the ®eld. The area of the sphere is
G A USS' S LAW 553

4pr 2 , and the total electric ¯ux out of the surface S of the sphere is
Z
q1 q1
E…r†  dS ˆ 2
 4r 2 ˆ : …15:14†
S 40 r 0
Equation (15.14) relates the ¯ux out of the sphere to the charge inside it.
This equation is Gauss's law, though here it has only been derived for the
very special case of a point charge at the centre of a sphere.

Surface integrals
In order to generalize Gauss's law to surfaces of any shape, we need to
work out the surface integral on the left-hand side of eqn (15.14).
Multidimensional integrals are discussed in Section 4.2 in Cartesian and
cylindrical polar coordinates. Here it is best to use spherical polar
coordinates, which are compared with Cartesian coordinates in Fig 15.9.
The position vector r of the point P has Cartesian coordinates (x, y, z).
The vector r can also be speci®ed by the spherical polar coordinates
(r, , ). The coordinate r is the length OP of r and  is the angle between Fig. 15.9 The spherical polar
coordinates …r, , † are de®ned
OP and the z-axis. The plane through OP and the z-axis cuts the xy-plane
with respect to a set of Cartesian
along OQ. The angle between OQ and the x-axis is the coordinate .
coordinates …x, y, z†.
The reason for using spherical polar coordinates here is that the proof
of Gauss's law depends on the mathematical concept of solid angle, which z Solid angles
is best expressed in this coordinate system. For readers unfamiliar with or
unsure of the meaning of solid angle, it is described in the box that
follows.

Solid angle is the measure of the angular size of a cone. Figure 15.10
shows part of a sphere with radius r and centre at the origin. The point P
with position vector r has spherical polar coordinates …r, , †. Keeping r
and  ®xed, rotate the position vector r through an angle . The point P
moves to Q along an arc of length r. Next rotate the line OQ through a
small angle  while keeping r and  ®xed at the values they have at Q.
The point Q moves to R along an arc of length r sin . When the D
rotations are performed in the order  followed by , P moves to R C
via S. Denoting the area of the spherical surface within PQRS by S, the
quantity

ˆ S=r 2 …15:15†

is called the solid angle subtended by the area PQRS at the origin O. The
area of the whole sphere is 4pr2 , and so S=4pr2 ˆ 
=4p is the Fig. 15.10 The area within PQRS is
fraction of the total area of the sphere covered by the area within PQRS. on the surface of a sphere of radius
Equivalently you may think of 
=4p as the fraction of the volume of r. The angles at the apex of the cone
the sphere occupied by the cone that has S as its base. Solid angle is are  and sin .
554 1 5 : ELE C TRO S TA TI CS

measured in the dimensionless units called steradians. The complete


sphere subtends a solid angle of 4p steradians at the origin.
Since the area PQRS is not ¯at, to calculate a solid angle we must
perform an integration. If  and  are made smaller and smaller,
z A sphere subtends a solid PQRS gets closer and closer to being a ¯at rectangle, and in the limit the
angle of 4p steradians at its in®nitesimal area dS ˆ r d  r sin  d and the in®nitesimal solid angle
centre of the cone with base dS is
d
ˆ dS=r 2 ˆ sin  d d: …15:16†
To ®nd the solid angle 
of the cone with angles  and  at the
apex, d
must be integrated over both  and ,
Z ‡ Z ‡

ˆ sin  d d:
 

To integrate over all directions, the limits are from  ˆ 0 to 2p, and
 ˆ 0 to p: if  were allowed to vary from 0 to 2p the whole sphere would
be covered twice. The total solid angle subtended by a sphere centred on
the origin is thus
Z 2p Z p
sin  d d ˆ 4p
ˆ0 ˆ0

con®rming, by direct integration, the result already derived from the area
of the sphere.

In Fig 15.11 a point charge q1 at the origin is surrounded by a closed


surface S. The cone with apex at the origin and solid angle d
ˆ
sin dd cuts through S at the point P with spherical polar coordinates
r, , and , and a small area dS of the surface lies within the cone. Because
the surface completely encloses the volume within it, a vector normal to
dS must point inwards or outwards. Gauss's law applies to the ¯ux of E
out of a closed surface, and the vector dS, of magnitude dS, is chosen to be
in the direction of the outward normal, as shown in Fig 15.11.
A sphere of radius r centred at the origin also passes through P, and
according to eqn (15.16) the area dSsphere of the sphere within the cone is
r 2 d
ˆ r 2 sin dd. The electric ®eld at P has a magnitude q1 =…4p0 r 2 †
and is perpendicular to the surface of the sphere. The ¯ux through dS is
the same as the ¯ux through dSsphere and is
Fig. 15.11 The ¯ux of E through q1 q1 q1
E…r†  dS ˆ  r 2 sin dd ˆ  sin dd ˆ d
,
the surface dS is determined by the 4p0 r 2 4p0 4p0
solid angle of the cone and the
and in the limit as dS becomes in®nitesimal
magnitude of the charge q1 , and
does not depend on the orientation q1
E…r†  dS ˆ d
: …15:17†
of dS. 4p0
G A USS' S LAW 555

The ¯ux of E through dS depends on the solid angle subtended by dS at


the charge q1 but not on the distance of dS or its angle to the position
vector r. The total ¯ux through S can now be evaluated using eqn
(15.17),
Z Z 2p Z p
q1 q1
Flux through S ˆ E  dS ˆ sin  d d ˆ :
S ˆ0 ˆ0 4p0 0
…15:18†

This result applies for any charge q1 and any surface S enclosing it. Any
number of charges qi within S will each give a contribution qi =0 to the
Fig. 15.12 The ¯ux through a
total ¯ux through S. closed surface due to a charge
There may in addition be charges outside S. Figure 15.12 shows that outside the surface is zero.
such charges make no contribution to the ¯ux over S. The cone from the
charge q1 passes twice through S, once entering and once leaving. The
®eld E entering S makes a negative contribution to the ¯ux out of S,
because the outward normal makes an angle greater than 90 with E at
this point. Where the cone leaves S the contribution is positive and, since
the solid angle is the same, the net ¯ux contributed by q1 is zero.
Figure 15.13 shows a more complicated surface S which is re-entrant. A
small cone with apex at a charge q1 within S must always pass outwards
through the surface one more time than it passes inwards, and the con-
tribution to the ¯ux is just q1 =0 as for a sphere. Similarly a cone from a
charge outside S may enter and leave more than once, but the number of
entering and leaving ¯uxes are the same, and the net contribution is zero.
The ®nal result, which is Gauss's law, is that, for any closed surface S,
Z P
qi Q
E  dS ˆ i ˆ …15:19† Fig. 15.13 Flux may pass several
S 0 0
P times in and out of a closed surface,
where Q ˆ i qi is the sum of all the charges situated within S. but for a charge located inside the
In words, Gauss's law states that surface the ¯ux always passes
outwards one more time than
The total ¯ux of the electric ®eld out of any closed surface equals the it passes inwards.
total charge enclosed within the surface divided by 0 .
Sometimes the electric ®eld possesses a symmetry that may greatly
simplify the calculation of the surface integral in Gauss's law. For
example, around a point charge q1 we can say immediately that the
electric ®eld must point towards or away from q1 and that its magnitude
must depend only on the distance from q1 . If we place q1 at the origin of
coordinates, the directions of the x-, y-, and z-axes are completely
arbitrary. One choice of directions for the axes is as good as any other. z When a system of charges
If we now use spherical polar coordinates related to the Cartesian possesses a simple symmetry,
coordinates as in Fig 15.10, all values of  and , which de®ne a the electric ®eld may be
direction, must be equivalent. The ®eld is said to possess spherical calculated easily using Gauss's
symmetry, and all points on a sphere centred at the origin are equivalent. law
556 1 5 : ELE C TRO S TA TI CS

There cannot be any component of the ®eld along the surface of the
sphere. The magnitude of E in the outward direction at a distance r from
the origin depends only on r and can be written as E…r†. According to
Gauss's theorem, the ¯ux through the sphere of radius r is E…r†
4pr 2 ˆ q1 =0 , and
q1
E…r† ˆ : …15:20†
4p0 r 2

The argument used to prove Gauss's law from the inverse square law has
been turned around by invoking the symmetry of the ®eld. Each law can
be derived from the other, and either may be used as the basis of
electrostatics.

Worked Example 15.4 A large number of small charges are placed close
together along a straight line so that the total charge per unit length is
 C mÿ1 (coulombs per metre). Assuming the line of charges to be
in®nitely long, calculate the electrostatic ®eld at a perpendicular distance
r from the line.
Answer This is an example of cylindrical symmetry, because all directions
pointing perpendicularly away from the line of charges are equivalent.
The electric ®eld must be perpendicular to the lineÐsince there is no way
to choose one direction along the line rather than the other, the
component of the ®eld along the line must be zero. A cylinder of
length ` with axis on the line and ends perpendicular to the
Fig. 15.14 The electric ®eld near a
line, as shown in Fig 15.14, is a suitable surface for the application of
line charge may be calculated by
Gauss's theorem. The ®eld is everywhere perpendicular to the curved
applying Gauss's law to an imaginary
surface of the cylinder and its magnitude E…r† depends only on the
cylinder of length ` and radius r.
distance r. The area of the curved surface of the cylinder is 2pr` and the
¯ux out of this surface is therefore 2pr`E…r†. The ¯ux out of the ends of
the cylinder is zero, since the ®eld lines do not cross the end surfaces. The
total amount of charge within the cylinder is ` and, applying Gauss's law
to the cylinder,
outward flux ˆ 2pr`E…r† ˆ total charge=0 ˆ `=0
or

E…r† ˆ : …15:21†
2p0 r

A real line of charges can never be in®nitely long. However, eqn (15.21) is
a good approximation for the magnitude of the ®eld provided that r is
small compared to the distance to the end of the line of charges.
T HE E LEC T R OST AT IC P OTE NT IA L 557

15.4 The electrostatic potential

The concepts of work and potential energy are discussed in general terms
in Sections 3.3 and 3.6. The work done by a force is de®ned in eqn (3.12)
as (force  the distance moved in the direction of the force). Examples
considered in Chapter 3 include the work done against the gravitational
force in lifting a mass, and against the restoring force of a spring when it
is stretched. In both cases energy must be expended to do the work, but
this energy does not disappear. It is stored as potential energy, which may
later be released: gravitational potential energy may, for example, be
released by allowing an object to fall.

Work done by electric charges


The concepts of work and potential energy also apply when the forces are
electrical. Consider the two positive charges q and q1 shown in Fig 15.15.
The charge q1 is ®xed, but q may be moved. There is a repulsive force
between the two charges and when they are separated by a distance r the
Fig. 15.15 The electric ®eld does
magnitude of the force is given by eqn (15.2) as
work on the charge q when it moves
qq1 from B to C.
Fˆ :
4p0 r 2
The force on q acts in the direction AB. If q moves a distance dr along the
line BC, the work done by the force is Fdr. The amount of work done
when q moves from B to C, changing the separation of the charges from
an initial value ri to a ®nal value rf is
Z rf Z rf  rf  
qq1 qq1 1 qq1 1 1
W ˆ Fdr ˆ 2
dr ˆ ÿ ˆ ÿ :
ri ri 4p0 r 4p0 r ri 4p0 ri rf
…15:22†

This work represents the difference in the electrical potential energy of


the two charges when q moves from B to C. It is natural to choose the
potential energy to be zero at rf ˆ 1, and with this choice the total
potential energy U of the two charges when they are at A and B, separated
by a distance ri , is
qq1
U ˆ W1 ˆ : …15:23†
4p0 ri
This equation is very similar to eqn (5.5), which gives the gravitational z Electrical and gravitational
potential energy of two masses, except that the sign is different. The sign potential energies are given by
change occurs because the gravitational force is attractive, whereas the similar expressions
electrical force between positive charges we have been considering here is
repulsive. Work must be done to pull the masses apart, and the
gravitational potential energy is therefore negative. The same applies to
558 1 5 : ELE C TRO S TA TI CS

charges of different sign. Equation (15.23) is still valid, but if q and q1


have different signs the right-hand side is negative, corresponding, as for
the gravitational case, to the fact that work must be done to pull the
charges apart.
The potential energy of q depends only on its distance from q1 and not
on the direction. In Fig 15.16 no work is done in moving q from B to E or
from C to D, since the electric force is perpendicular to the direction of
motion. Furthermore, the loss in potential energy in moving from B to C
may be recovered by using an external force to push q back to B. The
work done by the external force is also given by eqn (15.22). No work is
Fig. 15.16 The dashed lines are done if q moves from B to C and back again, nor is there any change in
perpendicular to the electric ®eld
potential energy. The same applies if q is taken round the path BCDEB or
due to the charge q1 and no work is
any other path starting and ®nishing at the same point: the potential
done if q follows the path BCDEB.
energy depends only on the position of q and not on the path it took to
get there. As explained in Section 3.6, a force that has the property of
z The electrostatic force is doing no work around a closed path is called a conservative force.
conservative Like the electric ®eld, the potential energy is usually most conveniently
expressed in terms of position vectors with respect to a ®xed origin.
Suppose that q and q1 are placed at points with position vectors r and r1
with respect to an origin at O, as in Fig 15.17,
The potential energy in eqn (15.23) may now be written
qq1
Uˆ : …15:24†
4p0 jr ÿ r1 j
If more charges q2 , q3 , . . . are now placed at r2 , r3 , . . . the potential
energy of q with respect to each one is an expression of the form of
eqn (15.24). The total potential energy, that is, the energy Utot that is
released if q is moved far away while all the other charges remain ®xed, is
Fig. 15.17 The potential energy of
q in the ®eld of q1 depends on the X qqj
Utot ˆ : …15:25†
distance jr ÿ r1 j between them.
j
4 0 jr ÿ rj j

The charge q has been used as a test charge to sample the potential energy
it gains in the neighbourhood of the ®xed charges qj . The potential energy
per unit charge Utot =q is determined only by the magnitudes and
positions of the ®xed charges qj , just as is the electric ®eld. The quantity
Utot =q is called the electrostatic potential and it is denoted by …r†,
X qj
…r† ˆ Utot =q ˆ : …15:26†
j
4 0 ÿ rj j
jr

Like the electric ®eld, the electrostatic potential is a function of


position. Unlike the electric ®eld, it has a magnitude but no direction: it is
a scalar ®eld. Since the potential represents energy per unit charge, it may
be measured in joules per coulomb. However, the potential is of such
great practical importance that it has a special unit called the volt,
T HE E LEC T R OST AT IC P OTE NT IA L 559

denoted by the symbol V. One volt is the same as one joule per coulomb.
One joule of energy is required to move a charge of one coulomb through
a potential difference of one volt.
The electrostatic potential depends linearly on the magnitudes of the
charges qj and, like the electric ®eld, the potential obeys the principle of
superposition. If the potential is known for two different sets of charges,
when both are present the potential is the sum of the potentials for each
separately.

Worked Example 15.5 The two electrons in a hydrogen molecule are


suddenly removed, leaving two protons separated by about 0.07 nm. The
protons then ¯y apart; calculate their ®nal kinetic energy.
Answer The potential energy of the two protons, given by eqn (15.18), is
converted entirely into kinetic energy. They have equal and opposite
momenta, and each has kinetic energy EK equal to half of the initial
potential energy,
e2 …1:602  10ÿ19 †2
EK ˆ ˆ  1:6  10ÿ18 J:
8p0 r 8p  8:854  10ÿ12  0:07  10ÿ9
The unit of energy on the atomic energy scale is the electron volt (eV),
which is the work done when a charge of e coulombs is moved through
a potential difference of one volt: 1 eV  1:602  10ÿ19 J. EK is thus
about 10 eV.

Equipotential surfaces
For an isolated charge q1 the electrostatic potential at a distance r from q1
is q1 =…4p0 r†. All points on the surface of a sphere of radius r are at the Fig. 15.18 Field lines radiating
outwards from the charge q1 are
same potential. Spherical surfaces centred on q1 are equipotential
perpendicular to the spherical
surfaces. No work is done in moving a test charge q across the surface
equipotential surface.
from one point to another. The electric ®eld generated by q1 points
radially outwards: electric ®eld lines pointing outwards from an
equipotential surface are illustrated in Fig 15.18.
At a point where a ®eld line crosses the equipotential surface the line is
perpendicular to the surface. This is obvious for a single charge, for which
the ®eld lines are radial and the equipotentials are spherical, but it is in
fact a general result. Electric ®eld lines are always perpendicular to
equipotential surfaces, no matter what the distribution of charge. This is
easily proved by considering a small movement of a test charge on an
equipotential surface. No work is done, and it follows that the electric
®eld does not have a component lying in the surface, that is, the ®eld is
perpendicular to the surface.
560 1 5 : ELE C TRO S TA TI CS

The electric ®eld and the electrostatic potential are really just different
ways of expressing the same information about a system of charges.
Figure 15.19 shows the ®eld lines and equipotentials in a plane passing
through a point charge q1 . Given a map of the contour lines representing
the equipotentials, we could draw ®eld lines cutting them perpendicu-
larly, and vice versa: given the ®eld lines, equipotentials cutting them at
right angles would have to be circles.
Up to now we have expressed the electric ®eld in units of newtons per
coulomb. Since one volt is the same as one joule per coulomb, newtons
per coulomb are equivalent to volts per metre (V mÿ1 ). Note that volts
per metre, which is the usual unit for describing electric ®eld, represents
the rate of change of potential with distance. In mathematical terms a
Fig. 15.19 Field lines (solid) and
rate of change is found by differentiation. For a point charge q1 ,
equipotential lines (dashed) in a
…r† ˆ q1 =…4p0 r†. The potential decreases with the distance r from
plane through a charge q1 .
the charge, and to ®nd the rate of change of potential with distance we
must differentiate with respect to r. Remembering that for a positive
charge the ®eld points outwards, in the direction of decreasing potential,
we have

d q1
ÿ ˆ ,
dr 4p0 r 2

the same as the magnitude of the electric ®eld of a point charge given in
eqn (15.20).
Another simple example relating ®eld and potential is illustrated in
Fig 15.20, which shows the ®eld lines for a uniform ®eld E pointing in the
z-direction. Two equipotential surfaces, which are both perpendicular to
the z-axis, are a distance d apart, and at potentials …z† and …z ‡ d†,
respectively. The force on a test charge q is qE and the potential energy at
P is q…z†. The difference in potential energy of the test charge between
Fig. 15.20 The relation between
the points P and Q equals the work done to move it back from Q to P,
®eld E and potential  is found by
moving a charge in the ®eld. The
q…z† ÿ q…z ‡ d† ˆ qV ˆ qEd: …15:27†
equipotentials are the dashed lines.

The difference in potential V between z and z ‡ d is usually called the


voltage difference or simply the voltage between the two points.
Remember that the ®eld points in the direction of decreasing potential.
In eqn (15.27) V and E are positive if …z† is greater than …z ‡ d†.
The uniform electric ®eld may also be represented in differential form
by allowing the distance d in eqn (15.27) to become very small. If d is
written as dz, then …z† ÿ …z ‡ dz† ˆ Edz and, in the limit as dz tends
to zero,

d
Eˆÿ : …15:28†
dz
T HE E LEC T R OST AT IC P OTE NT IA L 561

Notice that there is a minus sign in eqn (15.28), just as in the equation
relating ®eld and potential for a point charge.
A differential relation similar to eqn (15.28) applies to any electric ®eld and
the argument used here for the uniform ®eld is generalized in the box below.

Figure 15.21 shows two equipotential surfaces very close together, having
electrostatic potentials  and … ‡ d†. The vector d` is a vector in any
direction joining the point P on the surface at potential  to a point Q on
the surface at potential … ‡ d†. The electric ®eld E at P is perpendicular
to the equipotential surfaces. which are a distance PR = d` cos  apart.
The work done on a test charge q when it moves from P to Q is the
force qE times the distance PR in the direction of the force, i.e. Fig. 15.21 Here the charge is
moved in a direction different from
qEd` cos  ˆ qE  d`: This work is equal to the loss in potential energy
the direction of the ®eld.
…ÿqd†,
qE  d` ˆ ÿqd: …15:29†

The fact that no work is done on a test charge when it is moved round
any closed path that returns to its starting point is expressed mathe-
matically by taking the in®nitesimal limit of eqn (15.29) and integrating.
The result is that for electrostatic ®elds
I
E  d` ˆ 0: …15:30†
H
Here the symbol indicates that the line integral is round a closed path
made up of in®nitesimal segments d`.
If we choose Cartesian coordinates with unit vectors i, j, and k in the
x-, y-, and z-directions, we may write E ˆ E x i ‡ E y j ‡ E z k and d` ˆ
dxi ‡ dyj ‡ dzk leading to
E  d` ˆ Ex dx ‡ Ey dy ‡ Ez dz ˆ ÿd:
The partial derivative @=@x is the rate of change of  with x when
both y and z are kept constant. In the limit as dx, dy, and dz tend to zero,
@ @ @
Ex ˆ ÿ ; Ey ˆ ÿ ; Ez ˆ ÿ :
@x @y @z
The vector with components @=@x, @=@y, @=@z is called the gradient z The gradient of a scalar
of  and is written as grad . The three eqns above are summarized as function
E…r† ˆ ÿgrad…r†: …15:31†

The function grad…r† is a vector ®eld that has been derived from the
scalar ®eld …r†. The properties of grad have already been described
above in the discussion of the connection between the ®eld and potential:
grad is perpendicular to surfaces of constant  and its magnitude is the
rate of change of  with distance in that direction. In a two-dimensional
contour map the contours are equipotentials of the gravitational potential
562 1 5 : ELE C TRO S TA TI CS

 and grad  at a point on the map is in the direction of the steepest


uphill gradient from that point.

Exercise 15.4 An electron is placed in an electric ®eld of magnitude


100 V cmÿ1 . Calculate the electrostatic force on the electron and compare
it with the gravitational force.
Answer 1:6  10ÿ15 N. This is 1:8  1014 times greater than the gravita-
tional force on the electron.

The dipole potential and ®eld


Because the electrostatic potential is a scalar ®eld, it is often easier to
evaluate the potential than the ®eld for a particular distribution of
charges. Once the potential is known, the ®eld can be determined from
eqn (15.31). This is the method we shall use to calculate the ®eld in the
neighbourhood of an electric dipole consisting of two charges of equal
magnitude but opposite sign. The shape of the ®eld lines around a dipole
has already been sketched in Fig 15.6, but this ®gure is only a guess based
on the way the ®eld lines must pass from the positive to the negative
charge. The ®gure does not tell us how rapidly the strength of the ®eld
falls off with distance from the dipole, or precisely how it varies with
direction with respect to the axis of the dipole. It is important to have a
mathematical expression for the dipole ®eld, because it is responsible for
part of the interaction between molecules and it is also related to
electromagnetic radiation.
The dipole in Fig 15.22 has charges q separated by a distance a.
Spherical polar coordinates referred to a z-axis along the line joining the
charges are the most convenient for discussing the potential of the dipole.
The origin is midway between the two charges and the point P has
coordinates …r, , †: the symbol is used for the third coordinate in this
section, rather than the usual  to avoid confusion with the potential,
which is also usually labelled by . The dipole has cylindrical symmetry so
that all angles are equivalent and the potential depends only on r and .
The ®gure represents a plane passing through the z-axis and the point P.
The vectors r‡ and rÿ join the charges ‡q and ÿq to P. The potential at P
is the sum of the contributions from each charge, and is
 
q 1 1
…r† ˆ ÿ : …15:32†
4p0 r‡ rÿ
Fig. 15.22 The potential around a This expression applies everywhere, but it cannot be expressed simply
dipole is the sum of the potentials of in terms of the coordinates r and . In practice it turns out that what is
the two charges q separately. usually important is the ®eld at distances large compared with a.
T HE E LEC T R OST AT IC P OTE NT IA L 563

As the distance from the dipole to P increases, the fractional difference


between r‡ and rÿ becomes smaller. The two terms on the right-hand
side of eqn (15.32) therefore cancel each other more and more closely as
the distance r from the dipole to P increases, and the potential must
diminish faster than 1=r. When a=r is small, a good approximation to the
potential, which is worked out in the following box, is
qa cos  p cos 
…r† ˆ ˆ …15:33†
4p0 r 2 4p0 r 2
where p ˆ qa.

The magnitudes of the vectors r‡ and rÿ are given in terms of the


coordinates r and  by applying the cosine rule:
 
2 2 1 2 2 a2 a
r ˆ r ‡ 4a  ar cos  ˆ r 1 ‡ 2  cos  :
4r r
To ®nd the potential at distances large compared with a, we must
expand it in powers of a=r using the binomial theorem. All terms except z The expansion of 1=r in
the ®rst order in a=r will be neglected, and we may write polar coordinates
1 ÿ1=2 1 a ÿ1=2 1  a 
ˆ …r2 † ˆ 1  cos  ˆ 1  cos  :
r r r r 2r
Substituting these expressions for r in eqn (15.32), the leading terms in
1=r cancel and we are left with eqn (15.33).

The potential given in eqn (15.33) is often called the dipole potential, and
it represents the potential due to an idealized point dipole in which the
distance a has been allowed to tend to zero while p ˆ qa remains
nonzero. The dipole potential decreases with distance as 1=r 2 , whatever
the angle . As we predicted, this decrease is faster than 1=r because of the
cancellation of the ®rst-order contributions of the positive and negative
charges.
We have chosen the z-axis to lie along the line joining the charges of
the dipole. The vector p pointing along the z-axis and with magnitude p is z Expressing the dipole
called the dipole moment. In terms of this vector, the dipole potential is potential in terms of the dipole
pr moment
…r† ˆ : …15:34†
4p0 r 3
This expression makes no mention of the variable  and it is in fact
correct whatever may be the orientation of the dipole moment with
respect to the z-axis.
The general relation between the potential and the electric ®eld is E…r† ˆ
ÿgrad …r† (eqn (15.31)). In spherical polar coordinates the r-, -, and
components of E…r† are in the directions of the unit vectors labelled
564 1 5 : ELE C TRO S TA TI CS

er , e , and e in Fig 15.23. The component Er points directly away from


the origin. The component E is tangential to a circle passing through P
having constant r and , in the direction of increasing . Similarly, E is
tangential to a circle passing through P having constant r and , in the
direction of increasing . In terms of the potential, the components are
@ 1 @ 1 @
Er ˆ ÿ ; E ˆ ÿ ; E ˆÿ : …15:35†
@r r @ r sin  @
Here there is no variation with and the components are

p cos  p sin 
Er ˆ ; E ˆ ; E ˆ 0: …15:36†
2p0 r 3 4p0 r 3

The electric ®eld lines given by eqn (15.36) for small a=r are shown in
Fig 15.24. This ®gure applies to any plane that includes the z-axis. Both
the outward component Er of the electric ®eld and the component E
following circles of constant r and constant are proportional to 1=r 3 ,
falling off with distance faster than the ®eld due to a single charge. The
Fig. 15.23 The arrows show the terms in higher powers of a=r, which we have neglected, decrease faster
directions of the electric ®eld com- still. Electrically neutral molecules may possess a dipole moment and,
ponents E r , E  , and E at the point P although their dipole ®elds may cause important interactions with other
that has spherical polar coordinates molecules, the higher terms are almost always negligible.
…r, , †.

15.5 Electric ®elds in matter

Macroscopic electric ®elds


Inside a single atom the electric ®eld changes very rapidly with distance.
The atomic nucleus is extremely small, even on the atomic scale, and it
carries a charge Ze, where Z is the atomic number of the atom (which
determines the chemistry of the atom) and e is the electronic charge.
Close to the nucleus the positive charge on the nucleus is all that matters
and the ®eld is directed away from the nucleus. Further out, the negative
charge on the electrons tends to cancel out the effect of the nucleus, and
Fig. 15.24 The dipole ®eld due to outside the atom the ®eld is very small. On a microscopic scale these
a very small dipole with dipole changes of ®eld within an atom are extremely important, and indeed in
moment p. Chapter 11 the attraction given by Coulomb's law between an electron
and a proton is used to work out the properties of the hydrogen atom.
In this section we are concerned with the electrical behaviour of pieces
of matter made up of an emormous number of atoms. The ®elds within
individual atoms are not of interest: we need to know how the average
®eld varies over volumes large enough to contain very many atoms. Such
an average ®eld is called a macroscopic ®eld, to distinguish it from the
microscopic ®eld, which varies rapidly within atoms.
ELE C TRI C FIE LD S IN M AT TE R 565

Before discussing the macroscopic ®eld in an assembly of many atoms,


consider the average ®eld in a volume containing a single electrically
neutral atom such as the inert gas argon. The atom is spherically
symmetric, so that the ®eld within it is always pointing away from the
centre of the atom. The ®eld has a high value near the centre, like the ®eld
around a point charge, but it falls away even faster with distance because
of the negative charge on the electrons. To work out the average ®eld, you
have to remember that averaging a vector quantity is a bit different from
averaging a scalar quantity. Directions as well as magnitudes must be
taken into account. For a particular point with position vector r with
respect to an origin at the centre of the argon atom, the ®eld points in the
same direction as r. At the point diametrically opposite, which has
position vector ÿr, the ®eld has the same magnitude but is in the
opposite direction to the ®eld at r. The sum of the ®elds at r and ÿr is
zero. The same applies to all possible points r, and the average ®eld in a
volume including the atom is zero.
The example of an inert gas is a special case because the atoms are
spherically symmetric. However, except for some special materials, in the
absence of any electric ®eld applied from outside, the macroscopic electric
®eld in electrically neutral matter is zero. When charges are present, it is
not necessary to calculate the macroscopic ®eld by adding up the
contributions from every single particle carrying a charge e and then z The average ®eld due to
®nd the averageÐmost of the contributions just cancel out. The average electrically neutral atoms is
charge is determined within a volume small compared with everyday zero
objects, but still large enough to contain many atoms. The electric ®eld
caused by this average charge is then calculated.
Suppose that dVj is a small volume located at a point having a position
vector rj with respect to the origin. Let the net amount of charge within
dVj be …rj †dVj : …rj † is thus the charge density, that is, charge per unit
volume, measured in coulombs per cubic metre. Now divide up the whole
of the region containing charge into a lot of small volumes dVj . Each
contributes to the macroscopic electric ®eld, which by substitution in
eqn (15.11) is z Macroscopic ®elds are
calculated from average charge
1 X …rj †…r ÿ rj †dVj densities
E…r† ˆ : …15:37†
4p0 j jr ÿ rj j3

The average charge density …rj † varies smoothly with the position rj and
it is legitimate to replace the sum in eqn (15.37) with an integral, even
though we started with volumes dVj that are large enough to contain
many atoms. The macroscopic ®eld becomes
Z
1 …r0 †…r ÿ r0 †dV 0
E…r† ˆ …15:38†
4p0 volume jr ÿ r0 j3
566 1 5 : ELE C TRO S TA TI CS

where the integral, labelled volume, is over all volumes that contain a
net charge. From now on, when we refer to an electric ®eld E…r† without
stating whether it is a microscopic or macroscopic ®eld, we shall mean the
macroscopic ®eld that has been averaged over many atoms.
As already mentioned, the macroscopic electric ®eld in electrically
neutral matter is zero if there is no external electric ®eld. However, if an
object is placed in an electric ®eld, this ®eld is modi®ed by the presence
of the matter. To investigate how this comes about, we must consider
electrical conductors and insulators separately.

Conductors in electric ®elds


Materials like copper and aluminium that are good electrical conductors
are able to carry electric current because some of the electrons in the
material are free to move. These electrons, which are called conduction
electrons, are not ®xed to particular atoms, but are continually moving
through the material. In the absence of an electric ®eld there is no net
¯ow of charge, because the electrons are moving at random in all
directions. However, if a steady electric ®eld is applied, electrons, each
carrying a charge ÿe, experience a force in the opposite direction to the
®eld. There is a net ¯ow in this direction, and the ¯ow may continue for
an inde®nite time if the conductor is part of a complete electrical circuit.
On the other hand, if the conductor is isolated, the electrons cannot
continue to move when they reach the boundaries of the conductor. In
the slab of conductor shown in Fig 15.25, for example, the electric ®eld
pointing to the right causes electrons to migrate from the right-hand side
Fig. 15.25 Charges migrate to the to the left-hand side. Negatively charged electrons accumulate on the left-
surface of a conductor to ensure that hand surface, and the de®cit on the right-hand side causes a net positive
the electric ®eld is zero inside the charge to occur there.
conductor. The charges appearing on the surface of a conductor are called induced
charges. The induced charges themselves generate an electric ®eld
directed away from the positive charges towards the negative charges,
tending to cancel out the external ®eld. Conduction electrons will
continue to ¯ow, however small may be the resultant electric ®eld, and
they ¯ow until the electric ®eld within the conductor is zero. The
z The electrostatic ®eld inside disposition of surface charges depends on the shape of the conductor and
a conductor is zero is, in general, very dif®cult to work out. Whatever the shape, the charges
nevertheless arrange themselves so that the ®eld inside the conductor
is exactly zero. This applies to any material that contains conduction
electrons, and not just to very good conductors like copper. The
semiconductors silicon and germanium, for example, have conduction
electron densities billions of times smaller than copper at room
temperature but, when placed in a steady external ®eld, they also have
zero ®eld inside.
ELE C TRI C FIE LD S IN M AT TE R 567

Because the electric ®eld is zero throughout the conductor, its whole
volume is at the same potential. In particular, its surface is an equi-
potential surface. Since ®eld lines and equipotentials are always perpen-
dicular to one another, the external ®eld is normal to any conducting
surface. Using Gauss's law we can relate the magnitude of the electric ®eld
to the amount of charge on the conducting surface. In Fig 15.26 the
closed surface S is shaped like a pillbox. The curved surface is parallel to
the electric ®eld and there is no ¯ux through it. The ¯at surfaces of the
pillbox, each of area dS, are parallel to the conducting surface, one inside
and one outside the conductor. The electric ®eld and hence the ¯ux are
zero on the inside. The total ¯ux out of S is E  dS and, if the charge
inside the pillbox is dQ, Gauss's law gives
Z
E  dS ˆ EdS ˆ dQ=0
S

or
0 E ˆ dQ=dS ˆ  …15:39†

where  is the surface charge density, which is measured in coulombs per Fig. 15.26 Gauss's theorem relates
square metre (C mÿ2 ). In the simple example of slab geometry illustrated the induced surface charge to the
electric ®eld outside the conductor.
in Fig 15.26 the surface charge density is given directly in terms of the
external ®eld by eqn (15.39).
For other shapes of conductor the surface charge density must be
distributed in such a way as to ensure that the external ®eld is normal to
the conducting surface. This is illustrated schematically in Fig 15.27 which
shows a conducting sphere in an electric ®eld that is uniform far from the
sphere. Close to the sphere the surface charges modify the ®eld lines so
that they curve towards the sphere and meet it normally.

Exercise 15.5 The electric ®eld at the surface of a conductor is


104 V cmÿ1 . What is the surface charge density on the conductor, and
what average area has a charge equal to one electronic charge? Fig. 15.27 When a conducting
sphere is placed in an external
Answer The coulomb is a very large unit, and charge is frequently
electric ®eld E, the ®eld lines bend to
expressed in microcoulombs (1 mC  10ÿ6 C). The surface charge in
meet the conducting surface
this exercise is 8:85 mC mÿ2 . This is equivalent to one electronic charge normally.
on an area 1:8  10ÿ14 m2 or 1:8  104 nm2 , an area large enough to
accommodate about one million atoms.

The induced charges on the surface of a conductor are located in a very


thin layer. The amount of induced charge is given by the surface charge
density  and a surface integral must be added to eqn (15.38) to account
568 1 5 : ELE C TRO S TA TI CS

z Induced charges on the for the contribution of the induced charges to the ®eld. Including the
surfaces of conductors surface charges, the general expression for the macroscopic electric ®eld is
contribute to the electric ®eld Z Z
1 …r0 †…r ÿ r0 † dV 0 1 …r0 †…r ÿ r0 † dS 0
E…r† ˆ 3 ‡
40 volume
0
jr ÿ r j 4 0 surface jr ÿ r0 j3
…15:40†

where the labels volume and surface indicate that the volume integral is
over all volumes containing a volume charge density and the surface
integral is over all surfaces on which there is a surface charge density.
Similarly, the potential is
Z Z
1 …r0 †dV 0 1 …r0 †dS0
…r† ˆ 0j
‡ 0
: …15:41†
40 volume jr ÿ r 40 surface jr ÿ r j

Insulators in electric ®elds


In an insulator, all the electrons are ®xed to particular atoms. Over long
time periods, practically no migration of charge occurs when an
insulating material is placed in an electric ®eld. We can understand
how the material responds to the presence of a steady ®eld by considering
just one atom.
Imagine that a neutral atom is supported so that it does not fall under
gravity, but is free to move horizontally. If a horizontal electric ®eld is
switched on, there is no net force on the atom since its charge is zero.
However, the nucleus and the electrons experience forces in opposite
directions and they tend to move apart, without shifting the centre of
mass of the atom. As the centre of the distribution of negatively charged
electrons moves away from the positively charged nucleus, the mutual
attraction of nucleus and electrons creates a restoring force that balances
the force caused by the external ®eld.
Under all conditions that are met in the laboratory, the restoring force
is proportional to the distance x between the nucleus and the centre of the
electron distribution. Calling the constant of proportionality k, the
restoring force is kx. Figure 15.28 shows the forces acting on the nucleus,
but greatly exaggerates the relative movement of the electrons and the
nucleus: on the scale of the ®gure, the shift would not be visible for
Fig. 15.28 The force on the nucleus realistic electric ®elds. For an atom with atomic number Z and nuclear
of the atom due to the electric ®eld charge Ze, the force due to the external ®eld E is ZeE. This force is
E is balanced by a restoring force
balanced by the restoring force when kx ˆ ZeE, that is when x ˆ ZeE=k.
caused by the mutual attraction of
When the nucleus and the centre of electronic charge do not coincide,
the nucleus and the electrons.
the atom is said to be polarized. As for the point charges discussed in
Section 15.4, the vector in the x-direction and with magnitude equal to
ELE C TRI C FIE LD S IN M AT TE R 569

the product of the distance x and the charge Ze is called the dipole
moment of the atom and is measured in coulomb metres (C m). The
dipole moment is denoted by the vector p: the vectors x, p, and E all
point in the same direction and

…Ze†2
p ˆ Zex ˆ E ˆ 0 E …15:42†
k
where the constant is called the polarizability of the atom.
How does polarization affect the macroscopic electric ®eld in an
insulator? Let us ®rst consider a slab of uniform insulating material
placed in an electric ®eld normal to the faces of the slab. Within the slab
the macroscopic electric ®eld must be in the same direction as the ®eld
ouside the slab, and we shall assume for the moment that it is constant
throughout the slab, having a value Eint , say. Each atom of the insulator
acquires a dipole momen 0 Eint and, according to eqn (15.42), the centre
of the electron distribution is displaced a distance 0 Eint =Ze from the
nucleus.
The nucleus is much more massive than the electrons, and the centre of
mass almost coincides with the nucleus. We may picture the polarization
as if only the electrons move: this simpli®es the argument without
altering the results. Figure 15.29 represents a section through a slab of
insulator placed in an electric ®eld perpendicular to the sides of the slab.
The dashed lines show the boundaries of imaginary closed boxes with Fig. 15.29 The movement of
faces of area dS perpendicular to the ®eld: we shall apply Gauss's law to charge in the slab of insulator builds
these boxes. up charge on the surface but leaves
the interior electrically neutral.
When the atoms are polarized, electrons move through both surfaces of
the box (b), which is completely inside the insulator. Negative charge has
moved out of the left-hand side of the box, but just as much has moved in
at the right-hand side. The net charge inside the box is zero, as it was
before the atoms were polarized. Gauss's law tells us that the net ¯ux of E
out of the box is zero. This requires the ¯ux entering the left-hand side to
equal the ¯ux leaving the right-hand side, and the assumption that the
®eld is uniform within the slab is justi®ed.
Now look at the box (c), which straddles the right-hand surface of the
insulator. Negative charge has moved out of the left-hand side of the box
but there are no atoms at the right-hand side and the box has acquired a
net positive charge. If the number of atoms per unit volume is N , the
charge density of electrons is ÿNZe. All the electrons in the slab have
moved the same distance x, and the charge moving out of the area dS on
the left of the box is ÿNZexdS ˆ ÿNpdS. The box now encloses a net
charge ‡NpdS, and the surface charge density caused by polarization is
p ˆ Np. The opposite face of the slab acquires a surface charge density
(ÿNp† as electrons move into box (a). The slab as a whole is electrically
neutral, as it must be since it is composed entirely of neutral atoms.
570 1 5 : ELE C TRO S TA TI CS

If the surface of the insulator is at an angle  to the electric ®eld, as in


Fig 15.30, the surface charge density is reduced. If the area of the insulator
surface inside the box is dS, the projected area normal to the ®eld is
dS cos . The net charge inside the box is now p dS ˆ NpdS cos . At a
surface where the electric ®eld enters the insulator, p is given by the same
expression except for a minus sign. Remembering that the vector dS is the
outward normal to the surface, we ®nd that both the sign of the surface
charge density p and its angular dependence can be expressed concisely
by using the vector notation

p dS ˆ N p  dS ˆ P  dS …15:43†
Fig. 15.30 If the ®eld inside the
insulator is not perpendicular to its
where the vector P ˆ N p is the dipole moment per unit volume of the
surface, the charges move the same
insulator. The vector P is called the polarization density. The
distance, but the surface charge is
now spread out over a bigger area.
polarization density, like the dipole moment of a single atom, is in the
direction from negative to positive polarization surface charge.
The polarization density is useful because it is related to the electric
®eld inside the polarized material. In slab geometry this relation is easily
found from Gauss's law. The closed surface S in Fig 15.31 has surfaces of
area dS normal to the ®eld. The polarization charge within S is
p dS ˆ PdS. According to Gauss's law the net ¯ux out of S is therefore
PdS=0 . The ®eld entering S from the left is Eint and the ®eld leaving on
the right is Eext , and the net ¯ux is …Eext ÿ Eint †dS ˆ PdS=0 . Hence
Eext ˆ Eint ‡ P=0 …15:44†

or, since P ˆ Np ˆ N 0 Eint ,


Eext ˆ Eint …1 ‡ N † ˆ Eint …1 ‡ E †: …15:45†

The dimensionless constant E ˆ N is called the electric suscept-


ibility of the insulating material. The presence of the polarization charges
has reduced the ®eld inside the insulator by the factor (1 ‡ E ). This is
represented by drawing a reduced density of lines inside: in Fig 15.31
Fig. 15.31 The induced surface
some lines of the external ®eld end on negative polarization charges and
charge is related to the electric ®eld
inside and outside the insulator. start on positive polarization charges.

Worked Example 15.6 At 20 C and one atmosphere pressure helium gas
contains 2:7  1025 atoms mÿ3 , and the electric susceptibility of the gas is
6:5  10ÿ5 . Calculate the separation of the centres of the positive and
negative charges in a helium atom when it is placed in an electric ®eld of
106 V mÿ1 .
Answer From eqns (15.42) and (15.45) the separation is
0 E E 0 E
xˆ ˆ :
Ze ZeN
ELE C TRI C FIE LD S IN M AT TE R 571

Substituting the values given, the separation is z Polarization charge density


6:5  10ÿ5  8:85  10ÿ12  106
ˆ 6:7  10ÿ17 m ˆ 6:7  10ÿ8 nm,
2  1:6  10ÿ19  2:7  1025
a shift of about one-millionth of the radius of the helium atom.

Equation (15.45) has only been proved for slab geometry. The relation
between internal and external ®elds for other shapes of insulator is more
complicated, and we shall not discuss it in detail here. The direction of the
®eld as well as the magnitude may change at the boundary of an insulator.
Figure 15.32 shows the ®eld lines when an insulating sphere is placed in Fig. 15.32 The ®eld lines of an
a uniform external ®eld. The external ®eld lines are bent towards the insulating sphere placed in an
sphere, rather like the ®eld pattern for the conducting sphere. The sphere external electric ®eld E. The ®eld
is a specially simple case that can be solved exactly. The ®eld inside the changes direction at the surface
sphere is uniform: the ®eld inside has a smaller magnitude than the ®eld of the sphere.
outside because of polarization charges on the surface of the sphere.

Polarization charge density

Provided that an insulator is uniform, polarization polarization charge per unit volume, which we shall
charges appear only on the surface. For a non-uniform denote by p . Once the distribution of polarization
insulator there may be polarization charges distributed charges is known, the electric ®eld and potential are
throughout its volume. For example, if an insulator given by eqns (15.40) and (15.41), including polarization
consists of atoms that all have the same polarizability but charges, induced charges on conductors, and distribu-
a variable density, more charge moves in a more dense tions of free charge in the charge densities  and .
region than in a less dense one, and there is a net

Polar molecules
The atoms in a solid make small vibrations about ®xed positions. Each
atom is locked in place surrounded by neighbouring atoms, keeping the
same set of neighbours over long periods. When the solid is heated, the
vibrations become more and more energetic, until at the melting point
atoms escape from their ®xed positions and the solid turns into a liquid.
In many liquids the atoms do not move independently. They remain as
parts of a molecule with a ®xed structure. The molecules are the units that
change their positions and orientations with respect to their neighbours.
The water molecule, for example, consists of one oxygen atom and two
hydrogen atoms. As shown in Fig 15.33, the hydrogen and oxygen atoms
do not lie on a straight line. (The structure of the water molecule is brie¯y
explained in Section 11.4.) Furthermore, the oxygen atom has more than
its share of the electrons in the molecule, so that there is excess negative Fig. 15.33 The centres of positive
charge near the oxygen atom and excess positive charge near the and negative charge do not coincide
hydrogen atoms. The centre of positive charge does not coincide with the in the water molecule, and it has a
centre of negative charge, and the molecule has a dipole moment. dipole moment.
572 1 5 : ELE C TRO S TA TI CS

Molecules like water that possess a dipole moment are called polar
molecules. Although the water molecule is polar, in the absence of an
external electric ®eld the macroscopic electric ®eld inside a volume of
water is zero. This is because thermal motion ensures that all directions
are equally likely for the dipoles and, on average, their contributions to
the electric ®eld cancel out.
When a polar liquid or gas is placed in an electric ®eld, electrons and
nuclei are pushed in opposite directions just as in an insulating solid, and
as a result the liquid acquires a net dipole moment per unit volume.
There is an additional effect for polar molecules that is usually more
important. The dipole moments, which were initially randomly oriented,
are partially lined up by the ®eld so that they are more likely to be
pointing in the direction of the ®eld than opposite to it.
Fig. 15.34 The dipole placed in an
A polar molecule is represented in Fig 15.34 by positive and negative
electric ®eld experiences a couple.
point charges with the same dipole moment p ˆ qa as the moleculeÐthe
dipole moment is all that is needed for working out the effect of a
uniform electric ®eld acting on the molecule. The ®eld exerts a couple on
the molecule, tending to rotate it so that the dipole moment and the ®eld
point in the same direction. The ®gure shows that, when the dipole is at
an angle  to the ®eld, there is a couple qaE sin  ˆ pE sin  acting on the
dipole. This couple is zero for  ˆ 0 and for  ˆ 180 . At  ˆ 0 the
dipole is in stable equilibrium; when rotated through a small angle the
couple will turn the dipole back to  ˆ 0 . At  ˆ 180 the dipole is in a
position of unstable equilibrium; after a small de¯ection it will ¯ip over
to  ˆ 0 .
Work is done by the electric ®eld when it causes the dipole to rotate.
The potential energy of the dipole therefore depends on its orientation.
In Fig 15.35 equipotential surfaces passing through the charges ‡q and
ÿq are at potentials ‡ and ÿ respectively. From eqn (15.17) the
potential energy of the charge ‡q is ‡q‡ and, similarly, the potential
energy of ÿq is ÿqÿ . The potential energy of the dipole in the ®eld is
thus q…‡ ÿ ÿ †. The difference between the potential energies is given in
terms of the ®eld by eqn (15.29) as q…‡ ÿ ÿ † ˆ ÿE  a, leading to the
potential energy of the dipole
Fig. 15.35 The potential energy of Udipole …† ˆ q…‡ ÿ ÿ † ˆ ÿqE  a ˆ ÿp  E ˆ ÿpE cos : …15:46†
the dipole depends on its orienta-
tion in the ®eld. In calculating this potential we have not considered the energy of each
charge q due to the presence of the other. For a real molecule this
additional energy contributes to the binding energy of the molecule,
which does not change as the molecule is rotated in electric ®elds that can
z Thermal motion counteracts be realized in practice.
the tendency of dipoles to line If there were no thermal motion, all the dipoles in a polar liquid would
up along the electric ®eld line up with the ®eld. But the molecules are continually colliding with
their neighbours. There is a con¯ict between the thermal motion that
ELE C TRI C FIE LD S IN M AT TE R 573

tends to randomize the orientation of the molecules and the couple due
to the electric ®eld trying to line them up. The effect of thermal motion is
discussed in Section 12.6, where it is explained that the probability of
occurrence of states with different energies depends on the comparison of
the energy difference with a thermal energy kB T . Here kB is a universal
constant called Boltzmann's constant and T is the absolute temperature
measured in kelvins.
In a liquid the electric ®eld acting on each molecule is the internal ®eld
Eint , and the potential energy is ÿp  Eint . At ambient temperatures the
ratio pEint =kB T is always small, and the molecular dipoles have only a
slight tendency to line up with the ®eld. The molecular dipole moment
averaged over many molecules has a magnitude 13 …pEint =kB T †  p. The
average dipole moment is in the direction of the ®eld and, if there are N
molecules per unit volume, the polarization density P, which is the dipole
moment per unit volume, is
Np2
Pˆ Eint : …15:47†
3kB T
This result is proved in the box that follows.
The polarization density arising from the polarizability of the
molecules adds to that arising from the permanent dipole moment. For
an isotropic liquid or gas made up of molecules with a dipole moment of
magnitude p and polarizability , combining the results of eqns (15.45)
and (15.47) leads to an electric susceptibility
 
p2
E ˆ N ‡ : …15:48†
30 kB T

The probability of ®nding a molecule in a state with energy Udipole is given


by the Boltzmann factor exp…ÿUdipole =kB T † (expression (12.30)). Taking
the energy Udipole from eqn (15.46), the probability of ®nding a dipole
in a polar liquid at an orientation  to an external electric ®eld is
proportional to
   
ÿUdipole …† ÿpEint cos 
exp ˆ exp :
kB T kB T
When …pEint =kB T † is small, the exponential function may be expanded in
a Taylor series keeping only the ®rst term,
   
ÿpEint cos  pEint cos 
exp ˆ1ÿ :
kB T kB T
When the ®eld Eint is zero, all directions are equally probable, and the
probability of ®nding a dipole lying in the range of solid angle d
is
d
=4p ˆ sin dd=4p. Since all values of  are equally probable, only
574 1 5 : ELE C TRO S TA TI CS

the component p cos  of the dipole moment in the direction of Eint


contributes to the average dipole moment, which is
Z 2p Z p   
1 pEint cos  p2 Eint
p cos  1 ÿ sin  d d ˆ :
4p ˆ0 ˆ0 kB T 3kB T
If there are N molecules per unit volume, the polarization density P,
which is the dipole moment per unit volume, is
Np2
Pˆ Eint :
3kB T

15.6 Capacitors

Capacitors are used to store electric charge. They consist of a pair of


conductors with a potential difference maintained between them. Large
capacitors installed in oil-®lled tanks and operating with very high voltage
differenceÐup to many thousands of voltsÐmay accumulate large
amounts of electrical potential energy. At the other end of the scale
memory chips incorporate millions of tiny capacitors, each of which
represents the number 1 or 0 depending on whether they are charged or
uncharged. These memory capacitors operate with a few volts potential
difference between conductors separated by silicon oxide insulators.
Capacitors also have practical applications in electrical circuits carrying
currents that vary with time: this is discussed in Chapter 18.
The conductors in a capacitor are usually close together and separated
by a solid insulator. To study the properties of capacitors we shall start
by considering a parallel plate capacitor consisting of two parallel
conducting plates placed opposite to one another in vacuum as shown in
Fig 15.36. Suppose that there is a potential difference V between the
plates. In the diagram one plate is at earth potential ( ˆ 0) and the other
is at a positive potential V . This choice is only for de®niteness and, in
fact, the properties of the capacitor depend only on the potential
difference and not on the absolute values of the potentials on the plates.
The electric ®eld points in the direction of decreasing potential, and in
the centre of the capacitor the ®eld lines are straight from one plate to the
other. Near the edges of the plates the ®eld lines are still normal to the
conducting surfaces, which are equipotentials, but they bulge out as
shown and the electric ®eld does extend a little way outside the region
between the plates. However, these edge effects are rather small and, if, as
Fig. 15.36 Equal and opposite is usually the case, the separation of the plates is very small compared to
induced charges occur on the plates their length and width, it is a good approximation to assume that the ®eld
of the capacitor when they are held lines all pass straight across the gap between the plates and that the ®eld
at different potentials. outside is zero. The distance between the plates is d and, since their
C A PA CIT OR S 575

potential difference is V , the magnitude of the electric ®eld is E ˆ V =d,


from eqn (15.28). Electric ®eld lines start on positive charges, and positive
charges are induced on the plate at potential V . According to eqn (15.39)
the surface density of the induced charges is  ˆ 0 E ˆ 0 V =d. Negative
charges with the same magnitude of surface charge density are induced on
the plate at earth potential where the ®eld lines terminate.
For plates of area S, the total charge on the plate at potential V is
Q ˆ S ˆ 0 VS=d. Similarly, an induced charge ÿQ is located on the other
plate. The charges Q on the plates are proportional to the potential z The charge on the plates of
difference, and the proportionality constant is called the capacitance of a capacitor is proportional to
the capacitor. The capacitance is denoted by the symbol C, the voltage between them
Q ˆ CV …15:49†

where
0 S
Cˆ …15:50†
d
for a parallel plate capacitor in vacuum. The unit of capacitance is the
farad (symbol F). The farad, which has its magnitude ®xed by other SI
units, is impracticably large, and capacitances are usually quoted in
microfarads (1 mF  10ÿ6 F), nanofarads (1 nF  10ÿ9 F), or picofarads
(1 pF  10ÿ12 F).
The capacitance C is determined by the dimensions and geometrical
arrangement of the capacitor. Whatever the shape and size of two
conductors, it is always true that, when a potential difference is
maintained between them, equal and opposite charges are induced on
the conducting surface and the magnitude of the charge is proportional to
the potential difference. Equation (15.49) applies, with a value of the
capacitance determined by the geometry of the two conductors. Since the
charges on the conductors are equal and opposite, the total charge on any
capacitor is zero. The ¯ux of the electric ®eld through a surface enclosing
the capacitor is therefore zero and, apart from the small `fringing' ®elds
near the edges of the conductors, the ®eld outside the capacitor is
everywhere zero.

Relative permittivity
The capacitance of a parallel plate capacitor is given by eqn (15.50) only if
there is no matter between the plates, that is, the capacitor is in vacuum.
Figure 15.37(a) shows such a capacitor with charges Q on its plates. If
the capacitor is isolated, so that charge cannot ¯ow on to or away from Fig. 15.37 The capacitance is
the plates, the charge remains the same if a slab of insulator is placed increased by inserting dielectric
between the plates as in Fig 15.37(b). The slab is polarized and the electric material between the plates of the
®eld inside the insulator is less than the ®eld outside. Consequently, the capacitor.
576 1 5 : ELE C TRO S TA TI CS

Table 15.1 Relative permittivities of potential difference between the plates is also reduced, although
some materials, measured in steady the charge on them is unaltered. From eqn (15.45), the ®eld inside
®elds. The value for silicon is the insulator is smaller than the ®eld outside by the factor …1 ‡ E †. If the
included although it is too good a insulator ®lls the whole of the space between the plates, the ®eld has the
conductor to be used as the
reduced value everywhere between the plates, and the potential difference
dielectric material in a capacitor.
is also reduced by the factor …1 ‡ E †. The charges on the plates have
However, the relative permittivity of
semiconductors in steady ®elds has
remained the same, and the capacitance C ˆ Q=V (from eqn 15.44)) has
an important in¯uence on their increased by this factor,
behaviour capacitance with insulator between the plates
ˆ 1 ‡ E ˆ : …15:51†
capacitance in vacuum
Material Relative
permittivity 
The factor  by which the capacitance is increased by the insertion of
Mica 7.0 the insulating material is called the relative permittivity. An older name
Soda glass 7.5 for this factor is dielectric constant and, in the context of discussing the
Polyethylene 2.3 behaviour of electric ®elds, insulating materials are still usually referred to
Silicon oxide 3.9 as dielectric materials. The relative permittivities of some dielectric
Silicon 11.8 materials are given in Table 15.1.
The relative permittivity is simply a number for materials that are
Gas 104… ÿ 1) isotropic, that is, materials that have no directional properties. In some
Air 5.4 crystals induced dipole moments are not necessarily in the same direction
Ne 1.3 as the applied ®eld, but depend on the orientation of the ®eld to the
crystal axes. The equations we have derived are not then valid. Such
crystals have important applications in optics because of their effects on
the rapidly varying electric ®elds in visible light, but only isotropic
materials are used in capacitors. Provided that a capacitor is completely
®lled with a uniform dielectric material, the result that its capacitance is
enhanced by the factor  applies to all capacitors and not only to those
with slab geometry.

Worked Example 15.7 A coaxial cable consists of a wire with diameter


1 mm, passing through the centre of a polyethylene cylinder of diameter
3.5 mm, which is covered with a conducting coat made by braiding ®ne
wires. The outer conductor is held at earth potential. Calculate the
capacitance of a 1 m length of the cable.
Answer The cable has cylindrical symmetry and Gauss's law can be used
to determine how the ®eld varies within the cable. First, imagine that
there is no dielectric material between the conductors. Suppose that there
Fig. 15.38 (a) A coaxial cable con- is a positive charge  per unit length of the inner wire and a charge ÿ
sists of a wire through the centre of a per unit length on the outer conductor, as shown in Fig 15.38. If the
cylinder of insulating material within electric ®eld at a point at a distance r from the axis of the cable has a
an outer conductor. A cross-section magnitude E…r†, the ¯ux of E out of a cylinder of radius r and length `
through the cable is shown in (b). centred on the axis is 2pr`E…r† ˆ `=0 . Hence E…r† ˆ =…2p0 r†: The
C A PA CIT OR S 577

potential difference between points at r and r ‡ dr is d ˆ ÿE…r†dr from


eqn (15.29). The potential difference between radii a and b is
Z b Z b  
 dr  b
V ˆ d ˆ ÿ  ˆ ln ,
a a 2p0 r 2p0 a
and the capacitance of a 1 metre length of cable is
 2p0
Cˆ ˆ :
V ln…b=a†
When the space between the conductors is ®lled with polyethylene, the
capacitance per unit length is increased to 2p0 = ln…b=a†. The relative
permittivity of polyethylene is listed in Table 15.1 as 2.3. For a ˆ 0:5 mm
and b ˆ 1:75 mm, the capacitance of 1 metre is 1:0  10ÿ10 F or 100 pF.
Coaxial cables of this kind are frequently used to carry signals from one
piece of equipment to another, for example, from a receiving antenna to a
television set.

Stored energy
How much potential energy is stored in a capacitor when its plates have a
potential difference V ? We can work this out by starting with an
uncharged capacitor and gradually building up the charge, which is at all
times linked to the potential difference by eqn (15.49). Again suppose
that one plate is held at earth potential and the other carries a positive
charge as in Fig 15.36. When the positive charge has built up to a value
‡Q 0 the potential on the left-hand plate is V 0 ˆ Q0 =C. If further charge
dQ0 is moved from a great distance from the capacitor, where the
potential is zero, work V 0 dQ 0 ˆ Q 0 dQ 0 =C must be done. An extra charge
ÿdQ 0 is induced on the right-hand plate, but no work is required for this
since the negative charge does not change its potential. The total work
done in building up charges Q on the plates of an initially uncharged
capacitor, which is the potential energy stored by the capacitor, is
Z Q 0 0  
Q dQ 1 Q2 1
Uˆ ˆ ˆ CV 2 : …15:52†
0 C 2 C 2
For a 1 mC capacitor with 100 V across the plates, the stored energy is thus
1 ÿ6
2  10  104 ˆ 5  10ÿ3 J.

Exercise 15.6 A potential difference of 10 volts is maintained between the


two conductors of the coaxial cable in Worked example 15.7. Calculate
the energy stored in a 1 metre length.
Answer 5  10ÿ9 J.
578 1 5 : ELE C TRO S TA TI CS

Energy density of the electric ®eld


For a parallel plate capacitor with plates of area S separated by a distance
d in vacuum, the capacitance is given by eqn (15.50) as C ˆ 0 S=d. If the
potential difference between the plates is V , the energy stored by the
capacitor is 12 CV 2 (eqn (15.52)). We may equally well express this energy
in terms of the ®eld between the plates. Since V ˆ Ed, the energy can be
written as 12 C…Ed†2 ˆ 12 0 E 2 Sd. Now the volume of the capacitor is Sd
and we may think of the amount of stored energy as 12 0 E 2 per unit
volume. The same expression also applies to capacitors that do not have
slab geometry, and we can write the stored energy as
Z
2
Uˆ 1
2CV ˆ 1
2 0 E 2 dV …15:53†
V

where V is the volume where the electric ®eld due to the capacitor is
nonzero. For a steady ®eld it is not really possible to locate the energy
in a particular region of space and associating the energy with the ®eld
does not lead to any advantage in calculations. However, when there are
rapidly varying ®elds, as, for example, in the antenna of a mobile
telephone, energy is transmitted from one place to another by radiation.
It turns out that the energy density of the electric ®eld that we have
calculated for steady ®elds also applies to radiated energy. Similar
expressions, which we shall meet in the next chapter, apply to magnetic
®elds, and it is energy carried by both electric and magnetic ®elds that
constitutes electromagnetic radiation. The theory of radiation is beyond
the scope of this book, but it is important to realize that a thorough grasp
of the behaviour of electric and magnetic ®elds is required before
radiation can be understood.

Problems

Level 1 ®eld as a function of distance from the centre of the


sphere.
15.1 Calculate the force between two electrons that are
0.1 nm apart. 15.4 The sphere in Exercise 15.3 is hollow, having an
15.2 Two charges +q and one charge ÿq are placed at uncharged conducting concentric sphere of radius R1
the corners of an equilateral triangle of side a. What is inside the conductor. How much charge is there on the
the magnitude and direction of the force on the inner surface of the hollow sphere?
charge ÿq? 15.5 The maximum electric ®eld that can be sustained
15.3 A conducting sphere of radius R carries a total before breakdown in dry air is about 5  106 V mÿ1 .
charge Q. Draw a diagram showing the electric What is the minimum radius of curvature that can be
PR OB LE MS 579

tolerated on the corners of a box that is to be raised to has a lifting force of 1 N, and each has the same charge
100 kilovolts? Q. The angle between the strings is 30 . What is the
15.6 A charge q is at a distance d from a thin wire magnitude of Q?
carrying a charge  C mÿ1 . What is the force on the 15.14 A dipole consists of charges ‡q and ÿq separated
charge q? by a distance a. Derive an expression for the electric ®eld
15.7 Two parallel thin wires, each carrying a charge per on the axis of the dipole at a distance r from its centre, in
unit length  C mÿ1 , are separated by a distance d. What powers of …r=a†.
is the force per unit length between the wires? 15.15 Equal charges ‡q are situated at the corners of a
15.8 Two parallel wires, each of radius R, are a distance cube of side a. What force acts on any one of the charges,
d apart (d > 2R). One of the wires is at earth potential and what is its direction?
and the other at a potential V . Sketch the lines of the 15.16 A charge q is placed at the centre of a cube. What
electric ®eld and the equipotentials around the wires. On is the ¯ux of the electric ®eld through one of the cube
what part of the wires does the surface charge density faces? What is the ¯ux through one of the opposite faces
have its greatest value? if the charge is placed at a corner of the cube?
15.9 Two isolated plates are parallel to each other, One 15.17 A conducting sphere of radius a carries a charge
carries a total charge Q and the other a total charge 2Q. qa . Outside it are two thin conducting spherical shells, of
Use Gauss's law to ®nd out how the charges are distri- radii b and c…a < b < c† carrying charges qb and qc . The
buted on the surfaces of the plates, neglecting end effects. outermost sphere is at earth potential. Obtain expres-
15.10 A slab of material has a uniform charge density  sions for the potentials of the other two spheres.
throughout its volume. Calculate the electric ®eld as 15.18 Using the same value for the radius of the nucleus
a function of distance from the central plane of the slab. of the lead atom as in Problem 15.11, calculate the
15.11 The nucleus of a lead atom carries a charge 82e. It electrical potential energy of the nucleus. (Hint.
is quite a good approximation to assume that the Calculate the energy needed to build up the nucleus,
nucleus is a uniformly charged sphere of radius bringing charge from in®nity in in®nitesimal steps.)
7:5  10ÿ15 m. Draw a diagram showing the electric 15.19 Two dipoles, each with dipole moment
®eld as a function of distance from the centre of the 6  10ÿ30 C m, are placed as shown in Fig 15.40. Their
nucleus. What is its greatest value? separation a is 0.4 nm. Calculate the potential energy of
the dipoles. (These values apply roughly to water
Level 2 molecules.)

15.12 Three charges ÿq, ‡q, and ÿq lie on a line and


are separated by a distance a as shown in Fig 15.39. If
the negative charges are ®xed, calculate the restoring
force for small displacements, perpendicular to the line
Fig. 15.40
joining the charges, of the positive charge from its equi-
librium position half-way between the negative charges. 15.20 Calculate the energy of the dipoles in Problem
15.13 Two helium-®lled balloons are tied to the same 15.19 at the same separation, but when they are in line as
point on the ground by strings 50 cm long. Each balloon in Fig 15.41.
15.21 A capacitor in a random access memory consists
of a layer of SiO2 0.1 nm thick, sandwiched between

Fig. 15.39 Fig. 15.41


580 1 5 : ELE C TRO S TA TI CS

plane conductors each with an area of 0:6  10ÿ12 m2 . The three-dimensional equivalent of this problem
The relative permittivity of SiO2 is 3.9. Estimate the must be solved for different crystal structures to ®nd the
number of electrons on the negatively charged con- electrostatic contribution to the binding energy of ionic
ductor when the voltage across the capacitor plates is 5 V. crystals.
15.22 Two spherical conducting surfaces have radii a 15.28 Three concentric cylindrical conductors have
and b and the space between them is ®lled with air. radii 1 cm, 2 cm, and 3 cm. The space between them is
Calculate the capacitance of the two conductors. ®lled with oil with a relative permittivity  ˆ 2:2. If the
15.23 A conducting sphere of radius 1 cm is suspended maximum ®eld that the oil can maintain without
in air. Any other conductors are far away. Estimate the breakdown is 5  106 V mÿ1 , estimate the highest
capacitance of the sphere with respect to Earth. voltage difference that can be achieved between the
inner and outer conductors, and the voltage between the
15.24 A parallel plate capacitor has a capacitance of
inner and middle conductors under these conditions.
10 pF when the space between its plates is ®lled with air.
(The maximum voltage occurs when both the inner and
One of the plates is covered with a slab of dielectric
middle conductors have almost the breakdown ®eld on
material of relative permittivity 7, with a thickness that is
the outer surfaces.)
half the distance between the plates. What is now the
value of the capacitance? 15.29 A line charge of strength  C mÿ1 is parallel to an
earthed conducting plane and at a distance d from it,
as shown in Fig 15.43. Calculate the surface density
of the induced charge on the conducting plane as a
Level 3
function of y, neglecting end effects. (The solution of
15.25 Three charges ÿq, ‡2q, and ÿq are arranged on a this problem requires the use of the uniqueness theorem,
line as shown in Fig 15.42. Calculate the ®eld at a which states that if an electric ®eld is known to satisfy
distance r > a on the line, and ®nd the leading term in the conditions at the boundary of a region of space, it is
the expansion in powers of r=a. the only possible ®eld. Here the ®eld must be normal to
the conducting surface. Consider the ®eld due to two
line charges with strengths  and 2d apart. This ®eld
satis®es the boundary condition and in the region
between the line charge and the conductor it is the
required solution. The imaginary line charge ÿ is called
the image charge of the real line charge ‡.)
15.30 Two parallel plane conductors are a distance d
Fig. 15.42 apart. A plane sheet of charge with surface charge density

15.26 A slab of dielecric material 2 cm thick with


relative permittivity 4.0 has a net charge density
1 mC mÿ3 that is uniformly distributed throughout the
material. Calculate the electric ®eld inside and outside
the material.
15.27 A line of charges are all separated by a distance
0.1 nm from their nearest neighbours and the magnitude
of the charges is alternately ‡e and ÿe. Calculate the
potential energy of one of the charges due to the
interaction with all the others, assuming that there is a
very large number of charges. (Hint. It will help you to
®nd the answer to make a Taylor expansion of the
function ln…1 ‡ x†.) Fig. 15.43
AN SWE R S 581

 lies between the plates at a distance x from one of ®eld. The density of water is 0:9982 g cmÿ3 at 20 C and
them. Calculate the induced charge on each plate. 0:9718 g cmÿ3 at 80 C. Use these data to obtain the
15.31 The relative permittivity of water is 80.36 at 20 C dipole moment and the polarizability of a water
and 60.76 at 80 C when measured in a steady electric molecule.

Some solutions and answers to Chapter 15 problems

15.1 2:3  10ÿ8 N. plate carrying a total charge 2Q and ÿ 12 Q for the plate
15.2 Each of the charges ‡q exerts a force q2 =…4p0 a2 † carrying a total charge ÿ 12 Q.
on the charge ÿq. The horizontal components of these 15.10 The electric ®eld points outwards from the centre
forces are equal and opposite, and the net force on ÿq is and at a distance x from the centre, within the slab, its
down the page, as shown in Fig 15.44, with a magnitude magnitude is jxj=0 .
p 2 15.11 The ¯ux of the electric ®eld at a distance r from
2q2  3q
Fˆ cos 30 ˆ : the centre of the nucleus is due to the charge within r. If
4p0 a2 4p0 a2
r is less than the radius R of the nucleus, and the charge
15.5 Outside a conducting sphere of radius R the
density is , the total charge within a sphere of radius r is
potential varies as 1=r and may be written 4 3 2
3 pr . The area of this sphere is 4pr and the ¯ux out of
…r† ˆ …R†  R=r. The electric ®eld E…r† is
it is
@ R…R†
E…r† ˆ ÿ ˆ , 4pr 3  r
@r r2 4pr 2 E…r† ˆ , giving E…r† ˆ :
which has its maximum value …R†=R at r ˆ R. If the 30 30
corners of the box are rounded so that they are portions Outside the nucleus the electric ®eld is the same as for a
of spheres of radius R, the ®eld close to the corners is point charge at the centre, that is, it is proportional to
also …R†=R. The minimum radius Rmin that can be 1=r 2 and the maximum ®eld is at r ˆ R.
tolerated when the box is raised to a potential of 100 kV For lead the total charge 43 pR 3 ˆ 82e, and
satis®es
105 82e …82  1:6  1†ÿ19
ˆ 5  106 , leading to Rmin ˆ 0:02 m or 2 cm: Emax ˆ ˆ
Rmin 4p0 R 2 4p  8:85  10ÿ12  …7:5  10ÿ15 †2
15.7 The force is  2  1021 V mÿ1 :

2 This local ®eld that exists close to the nucleus is


N mÿ1 :
2p0 r enormous compared to the largest electric ®eld that can
15.9 There is a charge 32 Q on the outer surface of each occur even over distances about the size of an atom. As
plate. The charges on the inner surfaces are ‡ 12 Q for the indicated in Problem 15.5, the largest electric ®eld that is
sustainable in air is only about 5  106 V mÿ1 .
15.13 Assuming that the force between the balloons is
the same as if each were a point charge, the charge on
each balloon is 1:4 mC.
15.14 The ®eld on the axis is
 
qa 2a qa
3 1 ‡ 2 ‡  ˆ
4p0 jrj r 4p0 jrj3
for small …r=a†:
Fig. 15.44
582 1 5 : ELE C TRO S TA TI CS

The ®eld acts along the axis in the same direction, from 15.18 The electrostatic energy is calculated by building
the negative to the positive charge, on both sides of the up the charge on the nucleus from the centre, assembling
dipole. For small a=r the magnitude of the ®eld is thin spherical shells one by one like the successive layers
proportional to …1=r†3 : it falls off faster with distance of an onion. As in Problem 15.11, at a distance r from
than the ®eld due to a point charge, because to second the centre of the nucleus, less than its radius R, the
order in …a=r† the contributions of the positive and charge within r is 43 pr 3 . The potential at r is
negative charges cancel out.
1 4
15.15 The three charges at BDE and the three at CFH  pr 3 :
4p0 r 3
are symmetrically placed with respect to the diagonal GA
in Fig 15.45. By symmetry the force on the charge at A is
The energy needed to bring from in®nity an extra thin
therefore outwards in the direction GA from the
shell of radius dr, carrying a charge 4pr 2 dr is
opposite corner. The component of the force due to
the charge at B along GA is 1 4 4
p2 r 4 dr
 pr 3  4pr 2 dr ˆ 3 ,
4p0 r 3 0
q2 1
2
 p ,
4p0 a 3 and the electrostatic potential energy of the whole
nucleus is
the component due to the charge at C is
p Z R 4 p2 r 4 dr
4p2 R5 3 Q2
q2 2 3
ˆ ˆ 
p 2  p , 0 0 150 5 4p0 R
4p0 … 2a† 3

and the force due to the charge at G is where Q ˆ 43 pR 3 is the total charge of the nucleus. For
lead, Q ˆ 82e, giving a total electrostatic energy of
q2 1:24  10ÿ10 J or, equivalently, 775 MeV.
p :
4p0 … 3a†2 15.19 The two dipoles attract one another, and their
potential energy at a distance of 0.4 nm is ÿ0:126 eV.
The total force is therefore
15.21 The number of electrons is about 1300. Capaci-
 p 
q 2
3 3 2 1 q2 tors of about this size are used to store digits in dynamic
2
p 
 ‡  p ‡ ˆ 2:14 : random access memories (DRAMs).
4p0 a 3 2 3 3 4p0 a2
15.24 When no dielectric is present, the electric ®eld E
15.16 For a charge at the centre of the cube the ¯ux inside the capacitor is E ˆ V =d, where V is the voltage
of the electric ®eld out of one face is q=60 . For a difference between the plates and d their separation. If
charge at one corner the ¯ux out of an opposite face is their area is S, the charge Q on the plates is 0 ES and the
q=240 . capacitance C ˆ Q=V ˆ 0 S=d.
When the space between the plates is half-®lled
with an insulator, the ®eld inside the insulator is smaller
than the ®eld outside by a factor equal to the relative
permittivity . Call the electric ®eld at the surface of the
plates E 0 . The ®eld inside the insulator is E 0 = and
the voltage between the plates is V ˆ E 0 d=2 ‡ E 0 d=2.
The charge on the plates is now Q ˆ 0 E 0 and the
capacitance is

20 S 20 S
C0 ˆ ˆ :
Fig. 15.45 d ‡ d= d…1 ‡ †
AN SWE R S 583

The ratio of capacitances with and without insulator is

C0 2
ˆ :
C 1‡

For C ˆ 10 pF and  ˆ 7, the capacitance is increased to


17.5 pF.
15.25 The electric ®eld at a point P on the axis is
( )
q 1 2 1
Eˆ ÿ 2‡ 2ÿ
4p0 …r ÿ a† r …r ‡ a†2
(  )
q a2 ÿ2  a2 ÿ2
ˆ ÿ 1ÿ ‡2 ÿ 1 ÿ :
4p0 r 2 r r
Fig. 15.46

Making a binomial expansion in powers of a=r,


 surface. At a distance y from the line joining , the
q 2a …ÿ2†  …ÿ3† a2
Eˆ ÿ1 ÿ ÿ ÿ normal component is
4p0 r 2 r 12 r
a2 
2a 6qa2 2 d d
‡2 ÿ 1 ‡ ÿ 3 ‡ ˆ E? ˆ p  p ˆ :
r r 4p0 r 4 2p0 …y 2 ‡ d 2 † y 2 ‡ d 2 p0 …y 2 ‡ d 2 †
to second order in …a=r†. For small …a=r† the ®eld from
the charges falls off with distance as 1=r 4 , more rapidly The surface charge density  is related to the ®eld by
than the ®eld of the dipole in Problem 15.14. The E? ˆ =0 (eqn (15.39)), and the surface charge density
charges in this problem may be regarded as two dipoles on the conductor at a distance y from the line joining
close together, arranged so that their contributions the real and image charges is
almost cancel each other at large distances. Such an
arrangement of charges is called an electric quadrupole. d
ˆ :
15.27 The potential energy of one charge due to all the p…y 2 ‡ d 2 †
others is ÿe 2 ln 2=…2p0 d†. For d ˆ 0:1 nm this is
ÿ20:0 eV. 15.30 Let the potential at the plane occupied by the
positive charge density be . This plane is at a distance x
15.29 The potential due to the real charges ‡ per unit
from one of the conductors; choose the origin of x to be
length and the imagined `image' charges ÿ per unit
at this plane. The electric ®eld between 0 and x is ÿ=x,
length is zero everywhere on the plane midway between
and the induced charge density on the conductor at
them. The potential on the conducting plate is zero, and
x ˆ 0 is ÿ0 =x. Similarly, the induced charge density
the induced charge must be distributed on this plate in
on the conductor at x ˆ d is ÿ0 =…d ÿ x†. The total
such a way that the ®eld above the plate is the same as
induced charge density must be ÿ since there is no
the ®eld due to the charges  per unit length. Inside the
electric ®eld outside the conductors. Hence
conductor the electric ®eld is actually zero, and the
induced charges indeed move to the surface of the plane 0  0  x…x ÿ d†
to ensure that this is so, as illustrated in Fig 15.46. ‡ ˆ ÿ and 0  ˆ  :
x dÿx 2x ÿ d
The electric ®eld at the surface of a conductor is
always normal to the surface. To calculate the ®eld Hence the charge density on the conductor at x ˆ 0 is
due to the line charge and the image charge we only ÿ…x ÿ d†=…2x ÿ d† and the charge density on the
need to consider the component normal to the conductor at x ˆ d is ÿx=…2x ÿ d†.
584 1 5 : ELE C TRO S TA TI CS

For a single charge or any number of charges at a work by collecting such electrons and ions by placing
distance x from one of the conductors, the division of them in an electric ®eld. The electrons and ions move in
the induced charge between the two conductors is the opposite directions towards conducting plates. Induced
same as it is for a sheet of charge. X-rays and -rays charges appear on these conductors before the ions and
cause the formation of electrons and ions when they electrons arrive, allowing the time of generating the ions
interact with matter. Many X-ray and -ray detectors and electrons to be determined accurately.

You might also like