Vous êtes sur la page 1sur 65

Mechanics and Electromagnetism

D. Atkinson and O. Scholten



(August 1991)

These lecture notes were originally written by D.Atkinson for the courses of 1988/1989 and
1989/1990 and have been modied by O. Scholten for the course 1991/1992.
1
Chapter 1
Introduction
The purpose of this course is to explore mechanics and electromagnetism in the
formalisms of Lagrange and of Hamilton. The unication that this gives to the
subject is very satisfying; and the pivotal role that Special Relativity has in the
treatment of electromagnetic phenomena is especially clear.
It is assumed that the student has some knowledge of classical mechanics in the
formulation of Newton: from this springboard the method of Lagrange is developed
in Chapter 2, and that of Hamilton in Chapter 3. Two classical problems are handled
with the new tools in Chapter 4: the Kepler problem of planetary motion, that was
rst cracked by Newton, and the Rutherford scattering of -particles by an atomic
nucleus.
The reader probably knows the Lorentz transformation of Special Relativity
already: an interesting derivation is given in Chapter 5, and special importance
is attached to contravariant and covariant Lorentz vectors. The starting point of
Chapter 6 is the Maxwell system of electromagnetic equations, and these are cast
into manifestly relativistically covariant form. Lagrangian and Hamiltonian methods
are used consistently. Chapter 7 is devoted to the theorem of Emmy Noether, and
the conservation of electromagnetic energy, momentum, and angular momentum is
shown to follow from invariance of the lagrangian under respectively translations in
time and space, and under Lorentz transformations.
The eect of a eld on a point charged particle, and the inverse eect of such
a charged particle on the eld, are considered in Chapters 8 and 9. Some of the
technical details of calculations are relegated to the decent obscurity of a half dozen
appendices.
The main reference book used with the lectures is:
L.D. Landau and E.M. Lifshits, Mechanics and Electrodynamics, Pergamon Press.
In the notes at some places this book will be referred to. It is strongly recommended
to try to solve the problems given in the book. Other books that provide a more
extensive treatment of the material at approximately the same level are:
K.R. Symon, Mechanics, Addison-Wesley,
J.B. Marion, Classical Dynamics, Academic Press,
J.R. Reitz, F.J. Milford and R.W. Christy, Foundations of Electromagnetic Theory,
Addison-Wesey,
R.K. Wangsness, Electromagnetic elds, Wiley,
A much more indepth treatment of the material is given in the following two classics,
2
H. Goldstein, Classical Mechanics, Addison-Wesley,
J.D. Jackson, Classical Electrodynamics, Wiley.
3
Chapter 2
Euler-Lagrange Equations
2.1 Point Particle
AIM: derive equations of motions without the explicit introduction of forces.
REASON: Forces are often clumsy to deal with since it are vectors.
Consider a point-particle of mass,m, in a conservative force-eld: that is, suppose
that there exists a scalar potential energy, V (r ), that depends only on the position,
r , of the particle, such that the force is given by

F =

V. (2.1)
Generally, it is convenient to express vectors in terms of Cartesian components,
r = (x
1
, x
2
, x
3
) , thus:
F
i
=
V
x
i
. (2.2)
Newtons second law can then be written
d
dt
[m x
i
] =
V
x
i
, (2.3)
where dots mean time-derivatives.
In the alternative approach the non-relativistic kinetic energy of the particle is
introduced,
T =
1
2
m x
i
x
i
, (2.4)
the convention being made that a repeated index (here i) is to be summed from 1
to 3. Accordingly,
d
dt
[
T
x
i
] =
V
x
i
. (2.5)
The same equation (2.5) holds for N particles in a conservative force-eld. We
write the kinetic energy in the form
T =
1
2
m
i
x
i
x
i
, (2.6)
where we sum over the Cartesian coordinates of all N particles, with the convention
that m
3i2
= m
3i1
= m
3i
is the mass of the ith particle, the x, y, z coordinates
4
of which are x
3i2
,x
3i1
and x
3i
respectively. The potential energy may include
particle-particle interactions, so long as these are conservative:
V =
3N

i
V
i
+
1
2
3N

i=j
V
ij
. (2.7)
2.2 Canonical Transformation
AIM: rewriting equation (2.5) to obtain an equation of motion that is not only
valid for a cartesian coordinate system but for any set of generalized coordinates
describing the system.
We shall now perform a (possibly explicitly time-dependent) transformation of
the 3N Cartesian coordinates, x
i
, to a set of 3N independent coordinates, q
n
. A sim-
ple example would be the polar coordinates of the N particles, but much more exotic
possibilities exist, as we will see. The purpose is eventually to obtain coordinate-
independent equations of motion. By dierentiating Eq. (2.6) with respect to q
n
and to q
n
, we nd
T
q
n
= m
i
x
i
x
i
q
n
, (2.8)
T
q
n
= m
i
x
i
x
i
q
n
. (2.9)
Please note that in general
T
qn
,=
dT
dqn
and that care should be taken with the
dierence between straight and partial derivatives.
Since the q
n
are functions of the x
i
and t, but not explicitly of the q
n
, it follows
that x
i
may be written as a function only of the q
n
and t, so that
dx
i
dt
x
i
=
x
i
q
k
q
k
+
x
i
t
. (2.10)
Now the only dependence on the variables, q
k
, here is in the explicit occurrence of
q
k
on the right-hand side, so that
x
i
q
n
=
x
i
q
n
. (2.11)
Hence, combining eqs.(2.9,2.11),
T
q
n
= m
i
x
i
x
i
q
n
, (2.12)
so that
d
dt
[
T
q
n
] =
d
dt
[m
i
x
i
]
x
i
q
n
+ m
i
x
i
x
i
q
n
. (2.13)
By using Eq. (2.3), we can write the rst term on the right-hand side as

V
x
i
x
i
q
n
=
V
q
n
, (2.14)
5
since V does not depend on the x
i
explicitly. The second term on the right of
Eq. (2.13) is equal to T/q
n
, as can be seen from Eq. (2.8). Therefore
d
dt
[
T
q
n
] =
V
q
n
+
T
q
n
=
L
q
n
, (2.15)
where the Lagrangian is dened by L = T V . This is quite dierent from the
total energy, E = T + V . Since V does not depend explicitly on x
i
, neither does it
depend explicitly on q
n
, which means that the partial q
n
-derivative of T is equal to
that of L. Hence Eq. (2.15) can be rewritten as follows:
d
dt
[
L
q
n
]
L
q
n
= 0, (2.16)
which is the famous Euler-Lagrange equation, the derivation of which was the goal
of this chapter. The Lagrangian L should be regarded as a function of q, q, and t;
L(q, q, t).
2.3 Hamiltons Variational Principle
AIM: use a variational principle to arrive at the Euler-Lagrange equation and
introduce the principle of least action.
Finally, we introduce the action, as the time-integral of the Lagrangian:
S(t
2
, t
1
) =
_
t
2
t
1
dtL(q, q, t). (2.17)
Hamiltons variational principle states that, if we vary the functions q(t) and q(t),
for t
1
< t < t
2
, but in such a way that q(t
1
) and q(t
2
) remain xed at their physical
values, then, of all the possible values that q(t) and q(t) can have, those values that
make S extremal satisfy the Euler-Lagrange equations. We shall now prove this
statement.
S(t
2
, t
1
) =
_
t
2
t
1
dt[L(q + q, q + q, t) L(q, q, t)]
=
_
t
2
t
1
dt[
L
q
n
q
n
+
L
q
n
q
n
]. (2.18)
By an integration by parts, we nd
S(t
2
, t
1
) =
_
t
2
t
1
dt[
L
q
n

d
dt
(
L
q
n
)]q
n
, (2.19)
where the integrated terms vanish, since q is constrained to be zero at t
1
and t
2
.
Since q is arbitrary, it follows that for S to be zero and thus for S to be extremal,
the expression between square brackets must vanish.
There now appear to be two dierent axioms from which the same Euler-Lagrange
equation can be derived;
Newton: F = ma.
Hamilton: The action S is minimal.
Since these give rise to the same equation of motion, which determines the physical
observables, they are equivalent. For many generalizations used later, Hamiltons
principle is more versitile and it will therefore be used frequently.
6
Chapter 3
Hamiltons Equations
3.1 Canonical Momentum
AIM: introduce momenta associated with the generalized coordinates, the gen-
eralization of p
i
= m x
i
to non-cartesian coordinates.
In terms of Cartesian components, the Lagrangian of our system of N particles,
interacting with one another, and with an external, conservative eld, can be written
L(x, x, t) =
1
2
m
i
x
i
x
i
V (x, t), (3.1)
where V may depend explicitly on t (for example, a time-dependent external poten-
tial), but not on x. Then clearly
L
x
i
= m
(i)
x
(i)
, (3.2)
where there is no summation over the bracketed index, (i). We recognize the right-
hand side of Eq. (3.2) as a component of the linear momentum, and we now dene
the generalized, or canonical momentum, by
p
n
=
L
q
n
. (3.3)
It is important to realize that the canonical momenta are not necessarily momenta
in the Cartesian sense: for example, in terms of spherical polar coordinates, the
momentum conjugate to the angle, , is the angular momentum.
The rate of change of the Lagrangian can be written
dL
dt
=
L
q
n
q
n
+
L
q
n
q
n
+
L
t
. (3.4)
Now from the Euler-Lagrange equation, we know that
L
q
n
=
d
dt
[
L
q
n
], (3.5)
so that Eq. (3.4) can be rewritten
dL
dt
=
d
dt
[
L
q
n
q
n
] +
L
t
. (3.6)
7
If the potential and thus L has no explicit time dependence,
L
t
= 0, then from
Eq. (3.6) we obtain
d
dt
[
L
q
n
q
n
L] = 0. (3.7)
Since the time derivative of q
n
p
n
L is zero, it is associated with a conserved quantity
which is the total energy (see Eq. (3.10)) of the system.
3.2 The Hamiltonian
AIM: derive rst order equations of motion for the generalized coordinates and
the conjugate (canonical,generalized) momenta.
The Hamiltonian is dened by
H = p
n
q
n
L, (3.8)
and we use Eq. (3.3) to rewrite Eq. (3.6) in the form
dH
dt
+
L
t
= 0. (3.9)
If L does not depend explicitly on the time, the second term above is absent, and
hence the Hamiltonian is constant in time. If, moreover, the potential energy is
conservative, Eq. (3.2) holds,so that
H = m
n
x
n
x
n
L = 2T [T V ] = T + V, (3.10)
which means that the Hamiltonian is equal to the total energy, which is time-
independent. Note that two conditions are necessary for these conclusions to be
true: (1) L must not explicitly depend on the time, and (2) V must not explicitly
depend on the q.
From the denition, Eq. (3.8), we deduce
dH = q
n
dp
n
+ p
n
d q
n

L
q
n
dq
n

L
q
n
d q
n

L
t
dt. (3.11)
By using the denition (3.3), we see that the second and the fourth terms above
cancel. Moreover, the same denition, combined with the Euler-Lagrange equation,
implies that
L
q
n
= p
n
, (3.12)
so that
dH = q
n
dp
n
p
n
dq
n

L
t
dt. (3.13)
The form of this increment suggests that we consider H to be a function of q
n
, p
n
and t (and not independently of q
n
). Then the following partial derivatives can be
read o from Eq. (3.13):
q
n
=
H
p
n
(3.14)
8
p
n
=
H
q
n
(3.15)
L
t
=
H
t
=
dH
dt
. (3.16)
These are the Hamilton equations, which constitute an alternative to the Euler-
Lagrange system.
One of the advantages of Hamiltons approach is that one can readily dene more
general transformations of the variables. The set, q
n
,p
n
, may be replaced by another
set, Q
n
,P
n
, in which the Qs and the Ps can both be functions of all the qs and the
ps, and possibly of t. If the Hamilton equations, Eq. (3.14)Eq. (3.16), remain valid
in terms of the new variables (i.e. the transformation is such that these equations,
with q
n
and p
n
replaced respectively by Q
n
and P
n
,are true), then we speak of a
canonical transformation. The new variables are just as acceptable as the old ones.
P
n
is called the momentum canonically conjugate to Q
n
.
3.3 Conservation Laws
AIM: relate the existence of conservation laws to presence of symmetries (invari-
ants) in the system. Book: chapter 2
We have seen already that when L has no explicit time dependence i.e. L is
invariant under time translations, there is a conserved quantity, E, the total energy
of the system,
dE
dt
= 0 (see also book, 6 ).
It can also be shown (see book, 7) that if the system is invariant under transla-
tions, the operation r

i
= r
i
+ with being a constant (space and time independent)
vector, there is a conserved quantity, the momentum of the system. If the system
is invariant under rotations (see book 9) there is a conserved quantity, the angular
momentum.
In general it is the case that if the system has a symmetry, its properties are
invariant under a class of operations (space translations, rotations etc.) there will be
an associated conserved quantity. Finding the conserved quantities greatly facilitates
solving the equations of motion. for this reason it is important to nd the symmetries
of the system, the Lagrangian.
3.4 Poisson Brackets
AIM: derive the -so called- Poisson brackets which can be regarded as the classical
equivalent of quantum mechanical commutation relations.
Lastly, we introduce the Poisson bracket between two arbitrary functions of the
canonical variables ((q, p, t) and not q), say F and G:
F, G =
F
q
n
G
p
n

F
p
n
G
q
n
(3.17)
The time-derivative of the arbitrary function, F, can be written
dF
dt
=
F
q
n
q
n
+
F
p
n
p
n
+
F
t
9
=
F
q
n
H
p
n

F
p
n
H
q
n
+
F
t
= F, H +
F
t
(3.18)
In particular, the total time-derivative of a function that does not depend explic-
itly on the time is given by the Poisson bracket of that function with the Hamiltonian:
we say that the Hamiltonian generates time-translations. In particular,
q
n
= q
n
, H, (3.19)
p
n
= p
n
, H. (3.20)
These equations are impressively symmetrical between the qs and the ps, and they
replace the rst two of the Hamilton equations, (3.14) and (3.15). It is easy to see
that
q
k
, p
l
=
kl
, (3.21)
and Diracs recipe for the intuitive leap from classical to quantum mechanics is to
represent dynamical variables by linear operators on a Hilbert space, and Poisson
brackets by commutators (multiplied by 2/(ih)). With this interpretation, the
general equation (3.18), as well the quantization condition, Eq. (3.21), have the
same forms in classical and in quantum mechanics. Both Poisson brackets and
commutators satisfy the same algebra: in particular the Jacobi identity is satised
by both, namely
E, F, G +F, G, E +G, E, F = 0. (3.22)
10
Chapter 4
Central eld and Two-Body
Problems
4.1 One Dimensional Problem
AIM: show application of E-L equation of motion.
As a rst problem we will consider the case of a potential depending on a single
coordinate only, U(x),
L =
1
2
m x
2
U(x) (4.1)
Since there is no explicit time dependence the total energy E =
1
2
m x
2
+ U(x) is a
constant of motion. Using this we arrive at
x =
dx
dt
=
_
E U(x)/
_
1
2
m (4.2)
which is straightforward to integrate yielding
t =
_
1
2
m
_
dx
_
E U(x)
+ constant. (4.3)
As a specic example the harmonic oscillator potential U(x) =
1
2
kx
2
will be
taken, giving
t =
_
1
2
m
_
dx
_
E
1
2
kx
2
+ constant. (4.4)
With the standard substitution x =
_
2E/k sin the integration can be performed.
The full solution of the problem is thus
t =
_
m/k( +
0
),
x =
_
2E/k sin , (4.5)
or more familiarly,
x =
_
2E/k sin
_
k/mt +
0
, (4.6)
where
0
is determined by the initial conditions.
11
4.2 Two-Body Problem
AIM: show equivalence 2-body and central-eld problems (see book 11).
In this chapter, we shall consider a system of just two particles, at positions r
a
and r
b
, with an interaction potential energy, V , that depends only on the relative
positions of the particles. The Lagrangian can be written
L =
1
2
m
a

r
a
2
+
1
2
m
b

r
b
2
V (r
a
r
b
). (4.7)
We reduce this to an eective one-particle system by introducing the relative and
the centre-of-mass vectors:
r = r
a
r
b
, (4.8)
and

R =
m
a
r
a
+ m
b
r
b
M
, (4.9)
where
M = m
a
+ m
b
. (4.10)
In terms of these variables, the Lagrangian becomes
L =
1
2
M

R
2
+
1
2
m

r
2
V (r), (4.11)
where m is dened by
1
m

1
m
a
+
1
m
b
. (4.12)
The motion thus separates into a trivial one for the center of mass and one that is
more interesting for the relative coordinate. The latter is equivalent to that of a
particle of mass m moving in a potential V (r).
4.3 Central Field Problems
AIM: show the method of solution for a very general class of problems (see book
12).
A very common class of problems is that in which the force between two particles
depends on their relative distance only. This problem is equivalent to that of a
particle moving in a central eld, a potential V (r) depending only on r = [r[ and
not on its orientation, not on and in polar coordinates. In polar coordinates the
kinetic energy of a particle with mass m is
T =
1
2
m( r
2
+ r
2

2
+ r
2
sin
2

2
). (4.13)
The Lagrangian of the problem is
L = T V (r) =
1
2
m( r
2
+ r
2

2
+ r
2
sin
2

2
) V (r). (4.14)
12
The rst observation to make is that this L has no explicit dependence (is
cyclic in ) and thus there should be an associated conserved quantity. From the
Euler-Lagrange equation in the coordinate one deduces that
d
dt
L

=
L

= 0. (4.15)
The conserved quantity, a constant of motion, is therefore
M
z

L

=
1
2
mr
2
sin
2
2

, (4.16)
The z-component of the angular momentum.
By choosing a coordinate system in which initially = 0 and

= 0 one obtains
L
z
= 0 at t = 0 and at all later times since it it is a constant of motion. It is now
simple to solve Eq. (4.16), giving

(t) = 0 (4.17)
and thus (t) = 0.
Substituting Eq. (4.17) in the expression for the Lagrangian one observes that
(for this choice of coordinates) the lagrangian is cyclic in . From the Euler-Lagrange
equation one obtains the associated constant of motion,
M =
L

= mr
2

(4.18)
which is the angular momentum. From the conservation of angular momentum one
can easily derive Keplers second law (book 12). The value of M is known once the
initial conditions,

(t = 0) and r(t = 0) are specied.
Next the radial motion is solved, realizing that the problem is now reduced to
one dimension. The energy of the system equals
E =
1
2
m( r
2
+ r
2

2
) + V (r) =
1
2
m r
2
+
1
2
M
2
/mr
2
+ V (r) (4.19)
Using the approach discussed in the rst section one arrives at
t =
_
dr
_
2
m
(E V (r))
M
2
m
2
r
2
+ constant (4.20)
Alternatively the angular motion can be expressed as a function of r by starting
from
M = mr
2
d
dt
= mr
2
d
dr
dr
dt
, (4.21)
giving
d
dr
=
1
r
M/mr
2
. (4.22)
Integrating this gives the equation for the orbit of the particle,
(r) =
_
Mdr/r
2
_
2m(E V (r)) M
2
/r
2
+ constant. (4.23)
13
The integration constant is determined by the boundary condition at t = 0.
At the turning points the radial velocity r = 0. In general the kinetic energy
of the particle does not vanish since the angular velocity remains nite because of
angular momentum conservation. The position of the turning points can be obtained
from solving
E = M
2
/2mr
2
+ V (r). (4.24)
Please note that the direction of the angular motion (sign of

) remains xed.
4.3.1 Kepler Problem
A common problem (gravitation and coulomb) is one in which the potential is
proportional to 1/r, a problem investigated extensively by Kepler in the study of
planetary motion. The equation for the orbit of the particle, Eq. (4.23) can be solved
analytically using V (r) = /r ,
(r) = cos
1
_
_
M/r m/M
_
2mE + m
2

2
/M
2
_
_
+ constant (4.25)
In the book, 13, an extensive discussion of this problem and its special cases is
given and will not be repeated here.
14
4.3.2 Rutherford Scattering
book: 14,15,16
In the Kepler problem the interest was focussed on closed orbits or boundstates.
In Rutherford scattering one considers a charged probe coming in from a large
distance and scattering o a test charge (target). Since the size of each of the
particles is taken small compared to the typical size in the problem (the distance of
closest approach), the particles are taken to be point charces.
In a typical experiment one observes the scattered particles, in particular their
deviation in angle from a straight line trajectory. This deviation is the scattering
angle . In an imaginary experiment the scattering angle for each particle can be
determined with innite precision. In practice one aims a homogeneous beam of
particles with n particles per unit area, and measures the number of particles dN
scattered between the angles and

= + d. The partial cross-section for


scattering in this angle bin is dened as
d = dN/n. (4.26)
Please note that d carries the units of area.
To calculate this we use Eq. (4.23) to obtain
0
, the angle at which the projectile
reaches the distance of closest approach r
min
,

0
= (r
min
) =
_

r
min
Mdr/r
2
_
2m(E V (r)) M
2
/r
2
+ constant. (4.27)
where the constant in Eq. (4.23) is choosen such that the particle comes in at = 0.
Instead of using the energy and angular momentum as constants of motion it proves
to be easier to introduce instead the initial velocity of the particle,
E =
1
2
mv
2

, (4.28)
and the impact parameter ,
M = mv

. (4.29)
The scattering angle can be expressed in terms of
0
as
= [ 2
0
[, (4.30)
where time reversal symmetry has been used (incoming part of the orbit is same as
out going part). Since
0
depends on E and M or equivalently on and v

, also
the scattering angle can be regarded as a function of these two, (, v

). At this
point we can return to the problem of calculating the scattering cross-section.
Assume that a particle with impact parameter (

= + d) scatters at an
angle (

= +d), then dN equals the number of particles that were impinging


at the scatterer with impact-parameters between and

,
dN = n2[d[. (4.31)
The partial cross-section is
d = 2

d
d

d. (4.32)
15
As a last step the solid angle, d = 2 sin d is introduced to obtain the dierential
cross-section,
d
d
=

sin

d
d

(4.33)
which is a function of the scattering angle.
For the case of the coulomb potential, V (r) = /r the analytic solution can
be used for the orbits given in the previous section. After some manipulation the
details of which are given in 16 of the book, the famous Rutherford equation for
the dierential cross-section can be derived,
d
d
=
_

2mv
2

_
2
1
sin
4 1
2

. (4.34)
to describe the scattering of particles o a heavy nucleus.
The expression for the cross-section is the same for repulsive and attractive forces,
even though the orbits of the particles are very dierent.
The expression for the cross-section diverges for = 0, even the total cross-
section diverges. The reason is that even at large distances an 1/r potential is
non-negligible. The large impact parameters give rise to small angle scattering.
4.4 Small Oscillations
book 17,19,23
Many physical systems are in a quasi-stationary state in the sense that they
are close to their lowest energy state but have not reached it yet. The system can
be described in terms of small oscillations around this equilibrium state. Also for
investigating the stability of certain solutions one often considers the resonse of the
system under small oscillations.
Assume that a system has a equilibrium state for the coordimate q = q
0
. The
potential has a minimum at q
0
which will be normalized to zero, V (q
0
) = 0. Intro-
ducing the coordinate x = q q
0
, for small values of x the potential can be written
as
V (x) =
1
2
kx
2
+O(x
3
). (4.35)
In this expantion the linear term in x vanishes since the potential is required to have
a minimum at x = 0. Similarly the kinetic energy can be expressed as
T =
1
2
a(q) q
2
=
1
2
[a(q
0
) +O(x)] x
2

1
2
m x
2
. (4.36)
To leading order the Lagrangian is
L =
1
2
m x
2

1
2
kx
2
. (4.37)
The equation of motion is
m x + kx = 0, (4.38)
16
which has the well known solutions
x(t) = a cos(t + ),
=
_
k/m (4.39)
or equivalently
x(t) = '(Ae
it
),
A = ae
i
. (4.40)
From the Lagrangian the momentum can be constructed,
p =
L
x
= m x. (4.41)
As a next step the Hamiltonian is obtained,
H = p x L =
1
2
m x
2
+
1
2
kx
2
=
1
2
p
2
/m +
1
2
kx
2
, (4.42)
where the r.h.s. should be used for obtaining the Hamilton equations of motion. For
this system the energy, E =
1
2
m
2
a
2
is a conserved quantity, a constant of motion.
For a system with N degrees of freedom the approach is analogous, see book
19. The eigen frequencies (in general many, but not more than N) of the system
are obtained from solving the characteristic equation,
[k
ik

2
m
ik
[ = 0 (4.43)
with solutions

; = 1, , s ; s N. (4.44)
For each frequency there is at least one eigen-mode or normal-mode Q

in which the
system can oscillate with frequency =

. The general solution of the problem


can be expressed in terms of these normal-modes,
x
k
=

A
k
Q

. (4.45)
In problems like these it often will be advantageous to use the normal-modes as the
generalized coordinates of the system.
perturbations As mentioned in the begining of this section, the harmonic oscil-
lator Lagrangean represents only a lowest order (in the deviation from equilibrium)
approximation to the full problem. As an example of the use of perturbation theory
the following Lagrangean is considered,
L =
1
2
m x
2
+
1
2
m

x x
2

1
2
kx
2
(4.46)
with a small (m

m) anharmonic term in the kinetic energy. The equation of


motion for this problem is
m x + kx + m

(
1
2
x
2
+ x x) = 0. (4.47)
17
Since m

is small the solution is expanded in its powers, x = x


0
+m

x
1
+m
2
x
2
+ .
To zeroth order in m

the equation of motion is:


m x
0
+ kx
0
= 0, (4.48)
with the solution
x
0
(t) = A
0
e
it
;
2
= k/m. (4.49)
The part of the equation of motion proportial to m

reads
m x
1
+ kx
1
+
1
2
x
0
2
+ x x
0
= 0. (4.50)
Since the solution of the homogeneous equation is the solution for x
0
, only the special
solution needs to be obtained,
x
1
(t) =
1
2
A
2
0
/m[1 + e
2it
] . (4.51)
CHECK THIS EQN.
where the explicit form of the solution for x
0
(t), Eq. (4.49), has been used. This
procedure can be extended to higher order terms, but beware of some pittfalls, see
book 23.
18
Chapter 5
Special Relativity
5.1 Lorentz Transformation
The transformation between one coordinate system, (t, x, y, z), and another,
(t

, x

, y

, z

), such that the space axes of the two systems are coincident at t

= 0 = t,
and parallel thereafter, and such that the primed system has velocity v, along the
x-direction, with respect to the unprimed system, can be written
x

= (x vt) y

= y z

= z . (5.1)
The classical assumption (Galileo, Newton) was that = 1, but if we only require
that the origin of the primed system have velocity v with respect to the unprimed
system, the above more general transformation is possible. What is ? From the
uniformity of space, we see that must not depend on the coordinates (in a non-
uniform gravitational eld, near a star, for example, the properties of space are not
uniform, i.e. translation-invariant, but that is the domain of general, not of special
relativity). In the absence of gravity, is independent of t, x, y, z, but it can depend
on v. Space is also invariant under rotations: a rotation by 180 degrees about the
y-axis should have no measurable eect; but it changes the sign of the x and the
z coordinates, and also of the relative velocity of the primed system. Rotational
invariance therefore means that must be independent of the sign of v, i.e. it is a
function only of v
2
.
The unprimed system has a velocity v, with respect to the primed system,
along the x-direction, so the inverse relation between the two systems is
x = (x

+ vt

) y = y

z = z

, (5.2)
where is the same as in Eq. (5.1), since it is unaected by the change in the sign
of v.
Note that we have not assumed that the coordinate, t, in the unprimed system, is
the same as t

, the coordinate in the primed system. Such an assumption was made,


quite explicitly, by Newton, in his Principia: Absolute, true, and mathematical
time, of itself, and from its own nature, ows equably without relation to anything
external. If we set t

= t, with Newton, and do not enquire too closely what, if


anything, equable ow of time means, then Eq. (5.1) and Eq. (5.2) together imply
= 1. Now if light is emitted at the origin of space- and travels along the x-axis with
19
velocity c, then at t, the light will have reached the point x = ct. From Eq. (5.1),
with = 1, we see that x

= (c v)t, which means that the speed of light, as


measured in the primed system, should be c v.
The experimental fact that the measured speed of light is however independent
of the motion of the source, and of the observer (Michelson-Morley experiment and
the tests of ther-drag in the solar-system), forces the conclusions t

,= t and ,= 1.
Since the speed of light in the primed system is actually c, and not c v, it follows
that x = ct in the unprimed system must correspond to x

= ct

in the primed
system. Filling in these values into Eq. (5.1) and Eq. (5.2), we nd
ct

= (c v)t (5.3)
ct = (c + v)t

;
and by multiplying these two equations together, we obtain
= 1/

1
v
2
c
2
. (5.4)
By eliminating x

between Eq. (5.1) and Eq. (5.2), we see that


t

= (t
v
c
2
x) . (5.5)
From now on, we choose units of distance and time in such a way that c = 1, so
that we may rewrite the Lorentz transformation, as it is called, in the form
t

= (t vx) x

= (x vt) . (5.6)
Further, by introducing the parametrization v = tanh u, so that = cosh u and
v = sinh u, we can rewrite Eq. (5.6) in the form
t

= t cosh u x sinh u x

= t sinh u + x cosh u . (5.7)


This form looks very much like a rotation: if the space axes are rotated by an angle
about the z-axis, we have
y

= y cos x sin x

= y sin + x cos . (5.8)


There are two important dierences between Eq. (5.7) and Eq. (5.8), namely that,
in the former, hyperbolic functions replace circular ones, and the signs of the two
hyperbolic sines are the same, whereas those of the two circular sines are dierent.
One can regard the Lorentz transformation as a hyperbolic rotation between the
t-axis and the x-axis.
20
5.2 Contravariant and Covariant Vectors
AIM: introduce a compact notation without which formulation of Electromag-
netism will be very clumsy. The notation makes space-time equivalence manifest.
With c = 1, the relativistically invariant interval, ds, between two innitesimally
separated points is given by
ds
2
= dt
2
dx
2
dy
2
dz
2
(5.9)
In the following, we write x
0
= t, x
1
= x, x
2
= y, x
3
= z. In terms of the metric
tensor, g

, which is dened to be equal to be +1 if = 0 = , to be 1 if = i = ,


for i = 1, 2, or 3, and to be 0 if ,= , we can rewrite Eq. (5.9) compactly as follows:
ds
2
= g

. (5.10)
Here the Einstein summation convention has been used, that is, the repeated indices,
and , are implicitly summed from 0 to 3 . If a new set of coordinates, x

,are
linearly related to the old ones,
x

, (5.11)
where the matrix,

, is constant (i.e. independent of x), and is such that


g

= g

, (5.12)
then we speak of a Lorentz transformation. In Eq. (5.11) and Eq. (5.12), there is an
implicit summation over the repeated indices. This will be a general rule: if a Greek
index occurs above (a contravariant index), and below (a covariant index), in the
same term, then a summation over the values 0, 1, 2, 3 is implied. Any quadruple of
numbers, a

, together with the transformation law,


a

, (5.13)
is said to be a contravariant Lorentz four-vector.
Since the matrix,

, is independent of x, it follows, by dierentiation of


Eq. (5.11), that

=
x

, (5.14)
so that the transformation law, Eq. (5.13), can also be written
a

=
x

. (5.15)
Suppose now that is a Lorentz invariant function of x. Then, by the usual
chain rule for partial derivatives,

=
x

. (5.16)
Hence the partial dierentiation operator does not have the same transformation
law as the contravariant vector Eq. (5.15). Rather, it is an example of a covariant
vector, b

, which transforms as follows:


b

=
x

. (5.17)
21
The product of any contravariant vector, a

, and any covariant vector, b

, which we
often write simply as ab, is Lorentz invariant:
a

= a

=
x

= a

= ab. (5.18)
If a

is a contravariant vector, and we dene


a

= g

, (5.19)
then clearly
a

= g

= g

, (5.20)
where g

, the contravariant metric tensor, is equal to g

, the covariant tensor that


we introduced above. However, by dierentiating Eq. (5.12) with respect to x

, we
obtain
g

= x

. (5.21)
Next, dierentiate both sides with respect to x

, and recall that the matrix Eq. (5.14)


is independent of x:
g

= g

(5.22)
Multiply both sides by g

, and use the obvious fact that


g

, (5.23)
where the Kronecker is equal to 1 if both indices are equal, and to 0 if they are
not. Hence
g

=
x

. (5.24)
By using Eq. (5.20), we see that a

transforms as a covariant vector. In a similar


way, if b

is a covariant vector, then


b

= g

, (5.25)
is a contravariant vector.
22
CONTRA- & CO-variant Summarry
Take a point at time t and at space coordinates (x, y, z). In space-time coordi-
nates this point is specied as
Contravariant : x

= (t, x, y, z) , by denition. (5.26)


Covariant : x

= (t, x, y, z) , = 0, 1, 2, 3. (5.27)
The relation between co- and contra- variant vectors is thus given by
x
0
= x
0
x
i
= x
i
; i = 1, 2, 3. (5.28)
The invariant quantity is s
2
= t
2
x
2
y
2
z
2
. The metric tensor is thus given by
g

= g

=
_

_
1 = = 0
1 = = 1, 2, or 3
0 ,=
(5.29)
or
g =
_
_
_
_
_
1 0 0 0
0 1 0 0
0 0 1 0
0 0 0 1
_
_
_
_
_
. (5.30)
The metric tensor is obviously symmetric, g

= g

. The denition of a covariant


vector is
x

= g

(5.31)
with an implicit sum over repeated indices. Please check that using Eq. (5.26) this
leads to Eq. (5.27). By denition of the metric tensor the invariant length can be
written as
s
2
= t
2
x
2
y
2
z
2
= g

= x

= x

= x
nu
x
nu
= x
mu
x
mu
This last line is the reason for introducing the co- contra-variant notation. The
permutations and relabelling are allowed since we are not dealing with quantumme-
chanical operators but only with summations over real numbers, which commute.
In other words, a summation over individual matrix elements is implied, not the
multiplication of matrices. Always make sure that with a substitution of repeated
indices the correct number of implicit summations is implied;
(s
2
)
2
= x

(5.32)
with the implicit summation convention = 0, , 3 and = 0, , 3, the summa-
tion runs over 4 4 = 16 terms, is thus not equal to
x

(5.33)
since the summation convention is only = 0, , 3 and the implicit sum contains
a total of only 4 terms.
23
The contra/co variant notation has been introdeced because of the following rule:
RULE: If a quantity contains an equal number of upper (contravariant) and lower
(covariant) indices, and if each implicit summation runs over one upper and one
lower index and all indices are summed over in this manner, the quantity is an
invariant.
Please note that the introduction of upper and lower indices is only a matter of
bookeeping. It does not contain any magic, the above rule only ensures that the
correct metric tensors are included.
Lorentz transformations
Contravariant vectors transform as:
x

(5.34)
where

are the matrixelements of a 4 4 matrix, or, in sloppy language,

is
a 4 4 matrix. The matrix elements can be expressed as

=
x

(5.35)
Covariant vectors transform according to
x

(5.36)
where

= g

,=

(5.37)
24
5.3 Mechanics of a Free Particle
In order to guess the correct Lagrangian for a free particle in relativity theory,
and hence to infer the form of relativistic mechanics, we consider again Eq. (2.17),
which we can rewrite in the rest-frame of the particle, in which the -variable is ,
the proper of the particle, which is also the invariant interval:
(d)
2
= (dx
0
)
2
(dx
1
)
2
(dx
2
)
2
(dx
3
)
2
. (5.38)
In the rest-system, we write Eq. (2.17) in the form
S =
_
b
a
dL(x
i
, x
i
, ). (5.39)
The integration is taken from a space- point point, a, to another one, b. The problem
is that we do not know what L should be. It depends asymmetrically on space and
, and certainly is not a relativistic invariant. However, the action only refers to the
two space- points, a and b. This is true in all Lorentz frames, although of course
the and space coordinates of a and b do depend on the frame that is chosen. The
Hamilton variational principle simply says that the physical trajectory between a
and b is the one for which S = 0. This is a Lorentz invariant specication of the
dynamics, and it is consistent with the principle of relativity to require that the
action be a Lorentz invariant. However, the only invariants available, out of which
the action of a free particle could be made, are the mass, m, and the invariant
interval, . The action must be translationally invariant, so we cannot include
itself, but only the invariant measure, d. Hence Eq. (5.39) takes the form
S =
_
b
a
d, (5.40)
where is a function of the mass of the particle, m, only. Suppose that we now
transform from the rest-system to any other Lorentz frame. From Eq. (5.38), we see
that
d = dt
_
1 x
2
1
x
2
2
x
2
3
= dt

1 v
2
, (5.41)
where t x
0
, and v is the velocity of the new frame, with respect to the rest-
system. By comparing this equation with Eq. (2.17), we can read o the form of
the Lagrangian in a general Lorentz frame:
L =

1 v
2
. (5.42)
The constant, , can be identied by expanding the above expression to rst order
in v
2
:
L =
1
2
v
2
+ O(v
4
), (5.43)
from which it follows that = m, so that L reduces in the low velocity limit to
the correct non-relativistic kinetic energy,
T =
1
2
mv
2
, (5.44)
25
aside from the constant, m, which drops out of the Euler-Lagrange equation. The
Lagrangian is thus
L = m

1 v
2
. (5.45)
We next calculate the canonical momenta:
p
i
=
L
x
i
= m

x
i

1 x
j
x
j
=
m x
i

1 v
2
. (5.46)
The Hamiltonian is derived from the standard formula, Eq. (3.8), yielding
H = x
i
m x
i

1 v
2
+ m

1 v
2
=
m

1 v
2
. (5.47)
According to the general argument of Chapter 3, H will be a constant in time, and
we now assume that it is equal to the total energy, E, as in the non-relativistic case.
This assumption has far-reaching consequences, for we see that the energy of a free
particle in its rest-frame (v = 0), is not zero, but is equal rather to the rest mass
(in units for which c ,= 1, the rest-energy is mc
2
).
The form of Eq. (5.47) is not yet in canonical form, i.e. it is not expressed in
terms of the coordinates and momenta. To remedy this, we observe that
p
i
p
i
m
2
=
x
i
x
i
1 v
2
=
v
2
1 + 1
1 v
2
= A
2
1 , (5.48)
where A = 1/

1 v
2
. Hence
A =

1 +
p
i
p
i
m
2
; (5.49)
and by using the fact that H = mA, we obtain the Hamiltonian in canonical form:
H =
_
m
2
+ p
i
p
i
. (5.50)
Note that, for p
2
= p
i
p
i
<< m
2
,
H = m +
p
2
2m

p
4
8m
3
+ O(
p
6
m
5
). (5.51)
We recognise the second term as the nonrelativistic kinetic energy. The rst term,
the rest-mass of the free particle, is the equivalent energy that is locked up in
a particle at rest, and which can be liberated on the annihilation of matter and
antimatter.
26
5.4 Four-Momentum
Next, we introduce the 4-velocity,
u

=
dx

d
, (5.52)
which is clearly a contravariant 4-vector. From Eq. (5.41), we see that
u

=
1

1 v
2
dx

dx
0
(5.53)
so that
mu
0
=
m

1 v
2
= E, (5.54)
and
mu
i
=
m

1 v
2
dx
i
dx
0
= p
i
. (5.55)
Hence
p

mu

= (E, p ), (5.56)
is a contravariant 4-vector. The invariant,
p

= E
2
p
2
, (5.57)
has the same value in any Lorentz frame. In the rest-frame, it is clearly equal to
m
2
, so in general
E
2
= p
2
+ m
2
. (5.58)
As a simple application of this last formula, consider the decay at rest of a

+
meson, of mass m

, into a
+
lepton, of mass m

, and a neutrino, which has


mass zero. Since the 3-momentum is conserved, and it is zero before the decay, the
momenta of the
+
and of the neutrino must be equal and opposite. Suppose that
the magnitude of the momentum of the
+
, which can be measured, is p. The zeroth
component of the 4-momentum, the relativistic energy, is also conserved. Before the
decay, the energy of the
+
is just the pion mass (Eq. (5.58) with p = 0); and after
the decay, it is the sum of the the neutrino energy, which is equal to p (Eq. (5.58)
with m = 0), and the muon energy. That is,
m

= p +
_
p
2
+ m

2
. (5.59)
This equation can be solved to give
p =
m

2
m

2
2m

. (5.60)
27
Chapter 6
Maxwells Equations
AIM: Arrive at a lagrangian formulation of Electro-magnetism. This implies writing
the Maxwell equations as the equations of motion for the electromagnetic elds.
6.1 Electromagnetic Fields
AIM: write the Maxwell equations in a contravariant form.
The Maxwell equations, in the presence of a charge-density, (x), and a current-
density, (x), are

E = , (6.1)

B = 0 , (6.2)

E
t
= , (6.3)


E +

B
t
= 0 . (6.4)
Note that no polarization or magnetization has been included: these are the equa-
tions in vacuo, except in so far that charge distributions are taken into account. In
a polarizable and magnetizable medium, one distinguishes between

D and

E, and
between

B and

H. When, however, one adopts the more fundamental view that all
the charges should be explicitly taken into account, this distinction need no longer
be made. As in Chapter 5, we choose units such that c = 1 ; moreover the ugly
4 that often disgures the right-hand sides of Eq. (6.2) and Eq. (6.4) has been
removed by suitably redening the unit of electric charge.
Since

B is divergence-free (since there do not seem to be any magnetic monopoles),
it follows that there is a vector eld, A, whose curl it is, i.e.

B =


A. (6.5)
A proof of this statement can be found in the Appendix A, from which it will be
seen that

B does not determine

A uniquely. Dene

C =

E

A/t, so that


C = 0, (6.6)
28
and from Appendix A again, we know that this implies the existence of a scalar
eld, , such that

E


A
t
=

C =

. (6.7)
Substituting Eq. (6.5) and Eq. (6.7) into Eq. (6.2) and Eq. (6.4), we nd

t
2

2


t
[

t
+

.

A ] = ,

2

A
t
2

2

A +

[

t
+

.

A ] = . (6.8)
The above equations can be cast into a more compact form by writing the operator

2
/t
2

2
=

=
2
, and combining the scalar and the vector potentials into
one 4-potential:
A

= (,

A) . (6.9)
We shall show in a moment that this 4-potential has the transformation properties
of a contravariant Lorentz vector. We may write

2
A
0

0
[

] = , (6.10)

2

A +

[

] = .
These equations cry out to be combined into one covariant 4-dimensional equa-
tion, do they not? However, there is an awkward sign dierence in front of the
second term on the left. If we suppose that electric charge is a relativistic invariant,
so that the charge, e, of an electron is the same in any inertial system, then charge
density is not invariant; rather, the product, e = (x)d
3
x, is Lorentz invariant.
Electric current is caused by the ow of electrons: it is given by the sum of the
electric charges, multiplied by their velocities. The current density is accordingly
the product of the charge density and the velocity:
= v , (6.11)
so that if we dene the 4-current density by
j

=
dx

dt
(, ) , (6.12)
then
edx

= dx

d
3
x = j

d
4
x . (6.13)
Now since e and d
4
x are Lorentz invariants, and dx

is a contravariant 4-vector, it
follows that j

must also be a contravariant vector.


We can rewrite Eq. (6.11) in component form as follows:

2
A
0

0
[

] = j
0

2
A
k
+
k
[

] = j
k
. (6.14)
Now we can understand the apparently awkward sign dierence, for the derivative
operator is covariant, and if we cast it into the unnatural, contravariant form,

=
29

for = 1, 2, 3 , we pick up a minus sign! Hence Eq. (6.14) can be written in


the beautiful form

2
A

] = j

(6.15)
Since j

is a contravariant vector, it follows that A

must also be a contravariant


vector. In fact, Eq. (6.15), which is merely (!) a rewriting of Eq. (6.2)-Eq. (6.4),
is in a relativistically covariant form. The equations knew more than their creator,
Maxwell, did, when he invented them! To do Maxwell and Lorentz justice, they
were worried that the electromagnetic equations are not consistent with Galilean
covariance; and they did their best to understand this fact.
6.2 Electromagnetic Field Tensor
The contravariant 4-potential changes under a Lorentz transformation as follows:
A

(x

) =
x

(x) ; (6.16)
that is, the transformed eld, at the transformed point, is equal to the old eld, at
the old point, multiplied by the Lorentz-transformation matrix. A covariant version
of the 4-potential can be dened:
A

(x) = g

(x) ; (6.17)
and of course the transformation law for this is
A

(x

) =
x

(x) . (6.18)
It is convenient to introduce the second-order tensor
F

, (6.19)
which transforms as follows:
F

(x

) =
x

(x) . (6.20)
After these book-keeping preliminaries, we can write the Maxwell equations
Eq. (6.15) in the still compacter form

= j

. (6.21)
The eld tensor, which is manifestly antisymmetric, can be expressed directly in
terms of the electric eld and the magnetic induction, for if i,j and k are restricted
to the values 1, 2, 3, then
F
0k
=
0
A
k

k
A
0
=
0
A
k

k
A
0
= E
k
, (6.22)
and
F
jk
=
j
A
k

k
A
j
=
jkl
B
l
. (6.23)
30
Thus the eld tensor can be expressed wholly in terms of

E and

B, and vice-versa.
For future reference it is helpful to give the matrix elements F

in matrix form,
F


_
_
_
_
_
0 E
1
E
2
E
3
E
1
0 B
3
B
2
E
2
B
3
0 B
1
E
3
B
2
B
1
0
_
_
_
_
_
. (6.24)
Please note that F

is similar with some important sign changes,


F


_
_
_
_
_
0 E
1
E
2
E
3
E
1
0 B
3
B
2
E
2
B
3
0 B
1
E
3
B
2
B
1
0
_
_
_
_
_
. (6.25)
Also, since

E and

B are not four vectors but three vectors, E
1
= E
1
= E
x
.
Despite the fact that the 4-potential, A

, is not uniquely determined by the eld


tensor, it is an extremely useful quantity. If it is subjected to a gauge transformation,
i.e.
A

= A

G , (6.26)
where G is any Lorentz scalar eld, then clearly the eld tensor is unchanged. Such
a gauge-transformation has no physical consequences; and so any interactions with
the electromagnetic 4-potential must respect this gauge invariance. The restriction
turns out to be very important, with ramications far outside the eld of electro-
magnetism.
A particular restriction that is often made on the 4-potential is the so-called
Lorentz condition, viz.

= 0 . (6.27)
By means of a gauge transformation, it is always possible to achieve the Lorentz con-
dition, without changing the physics. For under the gauge transformation Eq. (6.26),

+
2
G . (6.28)
and the right side can be made to vanish by choosing G such that
2
G =

.
The solution of this partial dierential equation can be readily performed by Fourier
transformation, as in Appendix B.
When the Lorentz condition, Eq. (6.27), is satised, the Maxwell equations,
Eq. (6.21), become even simpler:

2
A

= j

. (6.29)
31
6.3 Lagrangian Density
Let us rst examine the free electromagnetic eld. We shall see how the Maxwell
equation, Eq. (6.21), can be derived from a variational principle. In order to do this,
we regard the eld, A

(t, r ), as a continuous innity of canonical variables. For a


given time, t, the canonical variables are labelled by , and the continuous variable,
r. Since the expression for the Lagrangian will inevitably involve a summation over
all space, it is convenient to introduce a Lagrangian density:
L =
_
d
3
xL(x). (6.30)
The action can accordingly be written
S =
_
t
b
ta
dtL =
_
t
b
ta
dt
_
d
3
xL(x) =
_
b
a
d
4
xL(x). (6.31)
Since the action, S, is a Lorentz invariant, and d
4
x is a Lorentz-invariant measure,
it follows that the Lagrangian density, L , is Lorentz-invariant.
Consider now a variation in the elds, A

, such that the values stay xed at the


space-time points a and b. The resultant change in the action is
S =
_
b
a
d
4
x[
L
A

+
L
(

)
(

)]
=
_
b
a
d
4
x[
L
A

L
(

)
]A

. (6.32)
Since A

is arbitrary between the end-points, the Hamilton variation principle,


S = 0, implies

L
(

)

L
A

= 0 (6.33)
This covariant expression is the continuum version of the Euler-Lagrange equation.
We shall now show that the Lagrangian density,
L =
1
4
F

=
1
2
F

, (6.34)
when inserted into Eq. (6.33), yields the Maxwell equation, Eq. (6.21). To do this,
we regard L as a function of the 4 variables A

, and the 16 variables

. We nd
L
A

= j

, (6.35)
and
L
(

)
=
1
2
F

=
1
2
F

]
= F

. (6.36)
It is clear now that the Euler-Lagrange eld equations yield indeed the Maxwell
equation, when the above Lagrangian density is used. However, the Lagrangian
32
density is not uniquely determined by the requirement that the equation of motion
be correct.
In the above account, the current density, j

, is simply treated as an externally


prescribed source of the electromagnetic eld. In a more thoroughgoing theory, this
source is itself expressed in terms of canonical elds, for example those pertaining
to the electron. This leads to quantum electrodynamics, which is a very beautiful
and successful theory, which is, however, beyond the present scope.
6.4 Hamiltonian
The canonical momentum densities that correspond to the eld variables, A

,
are dened by

=
L
(
0
A

)
= F
0
=
0
A

A
0
. (6.37)
From this, it is clear that
0
vanishes identically. Moreover, from Eq. (6.7), we see
that
E
i
=
0
A
i
+
i
A
0
= F
0i
=
i
. (6.38)
Remember that
i
=
i
, so that the signs above are correct. We have written the
components of the electric eld as subscripts:

E = (E
1
, E
2
, E
3
), and of course

E
and thus also is emphatically not a Lorentz vector.
The Hamiltonian density is now dened by
H =

L =
k

A
k
L
= F
0k

0
A
k
+
1
4
F

=
k
[
k
A
0

k
] +
1
4
F
j0
F
j0
+
1
4
F
0k
F
0k
+
1
4
F
jk
F
jk
=
1
2

k
A
0
+
1
4
F
jk
F
jk
, (6.39)
in current free space. The last form is canonical, that is, the hamiltonian density
is expressed in terms of the elds, and their space derivatives, but not their time
derivatives, and the momentum densities.
In terms of the electric eld and the magnetic induction, the hamiltonian may
be written in the form
H =
_
d
3
x H(x) =
_
d
3
x [

E.

A
0
+
1
2

E.

E +
1
2

B.

B] , (6.40)
which is not a lorentz scalar.
33
Chapter 7
Conservation Laws
7.1 Noether Theorem
THEOREM:
For every continuous symmetry of the Lagrangian density, there is a conserved
physical quantity.
We shall illustrate this theorem by considering the invariance of the free elec-
tromagnetic Lagrangian density, Eq. (6.34) without the current density term, under
time translations, space translations, and Lorentz transformations. These invari-
ances lead respectively to the conservation of energy, momentum, and angular mo-
mentum.
Suppose, in general, that L is unchanged under a transformation of the space-
time points x x

, and of the elds, A

(x) A

(x

). That is
L(

(x

)) = L(

(x)). (7.1)
Dene
A

(x) = A

(x) A

(x), (7.2)
and
L(x) = L(

(x)) L(

(x)). (7.3)
Note that A

(x) and

(x) occur in these denitions, and not A

(x

) and

(x

) . We nd therefore that
L =
L
A

(A

) +
L
(

)
(

), (7.4)
where the space-time argument, x, has been suppressed. Now

[
L
(

)
A

] =

[
L
(

)
]A

+
L
(

(A

),
=
L
A

+
L
(

)
(

), (7.5)
34
where the Euler-Lagrange equation, Eq. (6.33), has been used to obtain the last line.
On comparing this with Eq. (7.4), we see that
L =

[
L
(

)
A

], (7.6)
which is the general form of the Noether equation.
To evaluate L(x) in detail, we replace x

in Eq. (7.1) by x, which means that


we must replace x by x, where the transformation x x is the inverse of x x

.
So in place of Eq. (7.1), we have
L(

(x)) = L(

(x)). (7.7)
Thus Eq. (7.3) can be written
L(x) = L(

(x)) L(

(x))
= L(x) L(x)
= (x x)

L(x) + O((x x)
2
)
= x

L(x) . (7.8)
In the last line use has been made of the fact that the transformation is innetesimal
and continuous and thus (x x)

= (x x

.
7.2 Energy Momentum Tensor
Under space-time translations, x

= x

+a

, we know that the transformed


eld at the transformed point is just the original eld at the original point, i.e.
A

(x

) = A

(x), so that A

(x) = A

(x), where x

= x

. Thus
A

(x) = A

(x a) = A

(x) a

(x) + O(a
2
). (7.9)
Hence, for innitesimal a

,
A

(x) = a

(x), (7.10)
L(x) = a

L(x). (7.11)
The Noether equation, Eq. (7.6), takes the form
a

L = a

_
L
(

_
. (7.12)
Since only denotes an index for an implicit summation it may also be relabelled
by since this index has not been used on the right hand side. Bringing the terms
to one side and ipping positions of indices one obtains
0 = a

_
L
(

L
_
. (7.13)
Since this equation must hold for arbitrary a

, it follows that

= 0, (7.14)
35
where the energy-momentum tensor is dened by
T

=
L
(

L. (7.15)
By inserting the explicit form of the Lagrangian density for current free space,
Eq. (6.34), we obtain
T

= F

+
1
4
g

. (7.16)
The four-momentum (check that it is a contravariant vector indeed!) of the electro-
magnetic eld is dened by
P

=
_
d
3
xT
0
, (7.17)
so that, by using Eq. (7.14), we nd

=
_
d
3
x
0
T
0
=
_
d
3
x
i
T
i
= 0, (7.18)
on condition that the elds vanish at spatial innity. The four quantities, P

, are
the conserved quantities of the Noether theorem.
The zeroth component is just the Hamiltonian, since
P
0
=
_
d
3
xT
00
=
_
d
3
x[F
0

0
A

+
1
4
F

] =
_
d
3
xH = H, (7.19)
where we have used Eq. (6.39).
The space components of the eld four-momentum are
P
i
=
_
d
3
xT
0i
=
_
d
3
xF
0

i
A

=
_
d
3
x[
0
A
j

j
A
0
]
i
A
j
. (7.20)
However, from Appendix A, we see that
[

E

B]
i
= [

E (


A)]
i
= E
j
[
i
A
j

j
A
i
]
= [
0
A
j

j
A
0
][
i
A
j

j
A
i
]
= [
0
A
j

j
A
0
]
i
A
j

j
[
0
A
j

j
A
0
]A
i
. (7.21)
where we have used the free Maxwell equation to obtain the last line. Hence

P =
_
d
3
x

E

B, (7.22)
where

P = (P
1
, P
2
, P
3
). The integrand of the above equation, the eld momentum
density, is called the Poynting vector.
36
7.3 Angular Momentum Tensor
The Lagrangian density is invariant under under a Lorentz transformation; but
the four potential is not, since it is a four vector. If we write
x

, (7.23)
then
A

(x

) =

(x) . (7.24)
In order to obtain A, we need to calculate A

(x) rather than A

(x

). This is done
by replacing x

by x, which means that x must be replaced by x =


1
x :
A

(x) =

(x). (7.25)
We consider an innitesimal Lorentz transformation, and its inverse,

, (
1
)

. (7.26)
Hence
A

(x) = [

][1

]A

(x), (7.27)
so that
A

. (7.28)
Since the Lagrangian density is Lorentz invariant, we nd from Eq. (7.8) that
L(x) =

L(x) . (7.29)
By inserting Eq. (7.28) and Eq. (7.29) into Eq. (7.6), we obtain the following:

L =

[
L
(

)
A

+
L
(

)
x

]. (7.30)
Since

is antisymmetric (see appendix C), but otherwise arbitrary, it follows that


the odd part of its coecient in the above equation must vanish. Thus

= 0, (7.31)
where
L

=
L
(

)
A

+
L
(

)
x

L [ ]
= J

+ S

, (7.32)
where the extrinsic (or orbital) angular momentum density is
J

= x

[
L
(

L] [ ]
= T

, (7.33)
37
and the intrinsic (or spin) angular momentum density is
S

=
L
(

)
A

( )
= F

. (7.34)
We dene the angular momentum tensor by
L

=
_
d
3
xL
0
. (7.35)
In view of Eq. (7.31), we see that

=
_
d
3
x
0
L
0
=
_
d
3
x
i
L
i
= 0, (7.36)
so that all the components of the angular momentum tensor are time independent
(they are the Noether currents that correspond to the invariance of the Lagrangian
density under Lorentz transformation). Consider
J
2 1
=
_
d
3
xJ
0 2 1
=
_
d
3
x[x
1
T
0 2
x
2
T
0 1
] . (7.37)
Since T
0 i
are the momentum densities, we see that J
2 1
is the third component of
the orbital angular momentum, which we often write J
3
. Then S
3
S
2 1
is the
third component of the intrinsic, or spin angular momentum. We write L
2 1

L
3
= J
3
+ S
3
. In a similar way, by permuting 1, 2, 3 cyclically, we dene the other
components: L
1
= J
1
+ S
1
and L
2
= J
2
+ S
2
.
7.4 The Photon
We will now use the tensors, that we have introduced via the Noether theorem,
to calculate the energy, momentum and angular momentum in a particular eld
conguration, namely that specied by the following 4-potential:
A
1
= cos , A
2
= sin , A
0
= 0 = A
3
, (7.38)
where
= k

= k(x
0
x
3
). (7.39)
This corresponds to wave propagation along the x
3
axis, with unit velocity (i.e. the
speed of light). In terms of the electric and magnetic elds, and the eld tensor,we
nd
E
1
= B
2
= F
10
= F
13
= ksin (7.40)
E
2
= B
1
= F
20
= F
23
= kcos (7.41)
E
3
= B
3
= F
03
= F
12
= 0. (7.42)
It is easy to see that

= 0, so that the elds correspond to regions of space-time


in which there is no charge nor current density. Moreover, the electric and magnetic
38
elds are perpendicular to the direction of propagation, and to each other. Both
vectors have constant magnitude, but they rotate around the x
3
axis. We say that
the radiation is circularly polarized.
Consider rst the energy momentum tensor. We nd
T
00
=
1
2
[(F
01
)
2
+ (F
02
)
2
+ (F
23
)
2
+ (F
13
)
2
] = k
2

2
, (7.43)
so that the energy density is constant in space and time. Further,
T
0i
=
0
A
j

i
A
j
, (7.44)
so that T
01
= 0 = T
02
and
T
03
=
0
A
1

3
A
1

0
A
2

3
A
2
= k
2

2
, (7.45)
which is the same as the energy density.
Now suppose that a target is placed in this electromagnetic radiation, with cross-
sectional area orthogonal to x
3
, and suppose that it absorbs all the radiation that
falls on it, during a time T. The energy that is transferred to the target is obtained
by integrating T
00
over all the radiation that will fall on to the target, during the
specied period. This is, however, a volume of cross-sectional area and length T
(since the speed of light is unity). This absorbed energy is Tk
2

2
. Similarly, we
can calculate the momentum that is transferred to the target, in the same time, by
integrating T
0i
over the same volume. Clearly the momentum is purely in the x
3
direction, and it is also equal to Tk
2

2
.
In the 17th century, Newton postulated that light consists of a stream of particles.
This idea fell into disrepute, largely because of the success of the wave-hypothesis
of Huygens. The Maxwell equations are wave equations par excellence. At the
beginning of the 20th century, however, the particle theory of light was reintroduced
by Planck, in order to deal with the ultra-violet catastrophe of black-body radiation,
and by Einstein, in connection with the photo-electric eect, in which it was clear
that the particles of light, photons, each had an energy that was proportional to the
frequency.
Suppose that there are N photons, each of mass m, in the volume of the radiation
that falls on our target in the time, T. We calculated that both the energy and the
momentum of these N photons is Tk
2

2
. Hence the energy, E, and the momentum,
p, per photon are equal to one another. However, from Eq. (5.58), we know that
m
2
= E
2
p
2
, from which it follows that the mass of a photon is zero.
We will look lastly at the angular momentum transfer to the target. Since
T
01
= 0 = T
02
, it follows that J
3
vanishes. However,
S
021
= A
1
F
02
A
2
F
01
= k
2
. (7.46)
The total angular momentum that is contained in the volume of radiation, that falls
on the target in the time T, is therefore wholly intrinsic, and it is equal to Tk
2
.
The ratio of the energy to the angular momentum for the N photons, and thus for
each photon separately, is k = 2, where is the frequency of the radiation. Since
we know, by studying the energy of photo-electrons, that the energy of a photon is
proportional to the frequency, it follows that the intrinsic angular momentum, or
spin, of a photon is independent of the frequency. This spin does not in fact depend
on any of the accidental features of the eld conguration. It is a fundamental
unit of angular momentum, usually written as h/(2), where h is called Plancks
constant.
39
Chapter 8
Point Charge
AIM: solve the dynamics of Electromagnetic elds interacting with massive particles.
8.1 Lagrangian
The force on a particle of charge e, at rest in an electric eld

E, is e

E. The force
on a moving charge can be obtained by applying a Lorentz transformation to this
static situation, but some care is needed, since a three-dimensional force is not a
relativistic 4-vector. Consider the innitesimal quantity dened by
dp

eF

dx

, (8.1)
where e is the charge, and x the coordinate of an elementary charge, say an electron,
in an electromagnetic eld, F

. This is manifestly a Lorentz 4-vector. If we now


divide by dt, which is not a Lorentz scalar, but rather the zeroth component of a
4-vector, we obtain the non-covariant relation,
dp

dt
= eF

dx

dt
, (8.2)
the spacelike components of which (i.e. = j = 1, 2, 3) are
dp
j
dt
= eF
j0
+ eF
jk
v
k
, (8.3)
where v is the velocity of the electron. By means of Eq. (6.22) and Eq. (6.23), we
can write this, in 3-vector notation, as

p = e

E + ev

B . (8.4)
If the electron is instantaneously at rest, v = 0, then the right-hand side reduces to
e

E, the static force, which implies that p is the 3-momentum of the electron. We
know however that p constitutes the space components of the 4-momentum, which
is a contravariant 4-vector. Hence Eq. (8.1) is in fact the covariant relation between
an increment of the 4-momentum of an electron and the electromagnetic eld that
interacts with it.
To summarize the sequence of the reasoning, Eq. (8.1) is the denition of a
contravariant vector, which we identify with an increment of the 4-momentum of
40
the electron, for in the rest-frame of the electron the space components coincide
with the known static force exerted on the electron by the electric eld. Since
we know that the 4-momentum transforms under a Lorentz transformation as a
contravariant vector, we can assert that Eq. (8.1) identies the incremental change
of the 4-momentum in an arbitrary inertial frame. In a frame in which the electron
has velocity v , we see from Eq. (8.4) that the Lorentz force on a moving point charge
in an electromagnetic eld must be

F = e

E + ev

B. (8.5)
The zeroth component of Eq. (8.1), is also interesting, since it gives the incremental
change in the energy of the electron, E. Since F
00
= 0, the zeroth component of
Eq. (8.2) can be written
d
dt
E = eF
0k
dx
k
dt
= ev.

E . (8.6)
The rate of change of the energy of the electron is equal to the rate at which the
Lorentz force does work, which is

F.v, and so we see that Eq. (8.5) and Eq. (8.6)
are consistent with one another.
We shall now show that the Lorentz force, Eq. (8.5), can be derived from the
following action:
S = m
_
b
a
_
dx

dx

e
_
b
a
dx

. (8.7)
This can be rewritten
S = m
_
t
b
ta
dt

1 x
i
x
i
e
_
t
b
ta
dt[A
0
x
i
A
i
], (8.8)
where the indices of x and A are contravariant ones, and where the repeated Latin
index, i, is to be summed from 1 to 3.
From Eq. (8.8),we can read o the Lagrangian as
L = m

1 x
i
x
i
eA
0
+ e x
i
A
i
; (8.9)
and so the canonical momentum is
p
i

L
x
i
=
m x
i

1 v
2
+ eA
i
. (8.10)
Note that this canonical momentum, p
i
, is not the same as the ordinary momentum,
p
i
, the time derivative of which occurs in Eq. (8.4). Indeed p
i
= p
i
+eA
i
. L depends
explicitly on x
i
via the potentials A
0
and A
i
. Thus we calculate
L
x
i
= e
A
0
x
i
+ e x
j
A
j
x
i
. (8.11)
The Euler Lagrange equation (2.16) reads then
d
dt
[
m x
i

1 v
2
+ eA
i
] + e
A
0
x
i
e x
j
A
j
x
i
= 0. (8.12)
41
Now
dA
i
dt
=
A
i
t
+
A
i
x
j
x
j
, (8.13)
so it follows from Eq. (8.12) that
d
dt
[
m x
i

1 v
2
] = e
A
i
t
e
A
0
x
i
e
A
i
x
j
x
j
+ e
A
j
x
i
x
j
. (8.14)
Since
E
i
=
A
i
t

A
0
x
i
, (8.15)
and
(v

B)
i
= (v (


A))
i
=
A
j
x
i
x
j

A
i
x
j
x
j
, (8.16)
we see that the right-hand side of Eq. (8.14) is indeed the ith component of the
right-hand side of Eq. (8.5). This concludes the demonstration that the Lagrangian
(8.9) does indeed yield the Lorentz force.
8.2 Hamiltons Equations
It is convenient to make the transition from the Lagrangian to the Hamiltonian
formalism. The Hamiltonian is
H = p
i
x
i
L =
m x
i
x
i

1 v
2
+ eA
i
x
i
+ m

1 v
2
+ eA
0
e x
i
A
i
=
m

1 v
2
+ eA
0
. (8.17)
This is not yet in the canonical form, in which the Hamiltonian must be expressed as
a function of the coordinates, x
i
, and the canonical momenta, p
i
. From Eq. (8.10),
we have
m
2
+ ( p
i
eA
i
)( p
i
eA
i
) = m
2
+
m
2
v
2
1 v
2
=
m
2
1 v
2
, (8.18)
so that
H = eA
0
+
_
m
2
+ ( p
k
eA
k
)( p
k
eA
k
) , (8.19)
which is indeed canonical.
The Hamilton equations (3.8) take the form
x
i
=
H
p
i
=
p
i
eA
i
_
m
2
+ ( p
k
eA
k
)( p
k
eA
k
)
, (8.20)

p
i
=
H
x
i
= e
A
0
x
i
+
e( p
j
eA
j
)
_
m
2
+ ( p
k
eA
k
)( p
k
eA
k
)
A
j
x
i
. (8.21)
42
If the Lagrangian (8.9) does not depend explicitly on the time, the Hamiltonian
is independent of time: it is one of the constants of the motion. In such a case,
it is advantageous to use Eq. (8.19) to simplify the above two equations. From
Eq. (8.20), we have immediately
[H eA
0
] x
i
= p
i
eA
i
, (8.22)

p
i
+ e
A
0
x
i
= e x
j
A
j
x
i
. (8.23)
8.3 Constant Magnetic Field
In this section, we consider a charged point mass in a constant magnetic eld.
This is of relevance to the motion of elementary particles in a bubble chamber, or
in an accelerator. Suppose that A
0
= A
2
= A
3
= 0 but A
1
= Bx
2
, where B is
constant. Then all the components of

E and

B vanish, except B
3
=
2
A
1
= B.
The Hamilton equations, Eq. (8.22) and Eq. (8.23), are
H x
i
= p
i
+ eBx
2

i1
, (8.24)

p
i
= eB x
1

i2
. (8.25)
These equations reduce to
H x
1
= eB x
2
H x
2
= eB x
1
H x
3
= 0. (8.26)
The velocity in the x
3
direction is constant, and to solve the rst two equations, we
set = x
1
+ ix
2
, so that
H

= eB( x
2
i x
1
) = ieB

. (8.27)
The solution is
= + e
it
, (8.28)
where = eB/H. However, H = m/

1 v
2
, so that
=
eB

1 v
2
m
. (8.29)
From Eq. (8.28), we see that

= ie
it
; (8.30)
and hence
v
2

( x
1
)
2
+ ( x
2
)
2
= [

[
2
=
2
[[
2
. (8.31)
Thus v

is constant, and since


[ [
2
= [[
2
=
v
2

2
, (8.32)
43
it follows that the motion in the (1, 2) plane is a circle, of radius R = v

/. In view
of the constant velocity in the x
3
direction, it follows then that the trajectory of the
particle is a spiral, or helix, around the direction of the constant magnetic eld.
In the case that x
3
= 0, i.e. the motion is perpendicular to

B, so that v

= v,
the particle travels with constant speed in a circle of radius
R =
vm
eB

1 v
2
. (8.33)
In a synchrotron, in order to keep the radius of the trajectory constant, one has
to increase the bending magnetic eld, as the speed of the particle gets closer and
closer to that of light.
8.4 Constant Electric Field
Consider next a point charge in a constant electric eld, E. For convenience, we
choose the x
3
axis to be parallel to the eld, and the origin to be such that x
i
= 0
when t = 0, i = 1, 2, 3. Let the initial value of the particle velocity be (v
1
, v
2
, v
3
),
and choose the x
2
axis such that v
2
= 0. The elds can be specied by

A = 0, and
= A
0
= Ex
3
. The Hamilton equations of motion are
(H + eEx
3
) x
i
= p
i
, (8.34)

p
i
= eE
i3
. (8.35)
We can integrate Eq. (8.35) immediately and then combine it with Eq. (8.34):
(H + eEx
3
) x
i
= Hv
i
+ eEt
i3
, (8.36)
and the Hamiltonian,
H =
m

1 v
2
eEx
3
, (8.37)
is a constant.
For i = 2 this implies that x
2
0, i.e. that the motion is wholly in the (x
1
x
3
)
plane. For i = 3, we rewrite Eq. (8.36) as follows:
d
dt
[(H + eEx
3
)
2
] =
d
dt
[(Hv
3
+ eEt)
2
], (8.38)
from which we deduce
(H + eEx
3
)
2
(Hv
3
+ eEt)
2
= H
2
[1 v
2
3
]. (8.39)
By dividing the i = 1 and the i = 3 components of Eq. (8.34) , we obtain
dx
3
dx
1
=
Hv
3
+ eEt
Hv
1
. (8.40)
On using Eq. (8.39) to eliminate t in favour of x
3
, we nd the solution
1 +
eEx
3
H
=
_
1 v
2
3
cosh[
eEx
1
Hv
1
], (8.41)
44
which is a catenary, the same gure as that assumed by a uniform cord, hanging in
a uniform gravitational eld.
In the above work, we have relied heavily on the fact that

H = 0. The actual
value of the Hamiltonian can be found by computing Eq. (8.17) at t = 0:
H =
m
_
1 v
2
1
v
2
3
, (8.42)
since at t = 0, x
3
= 0, so that then also A
0
= 0. It is important to realise that
the velocities do change, and that although H is indeed time-independent, the two
terms in the last line of Eq. (8.17) separately do depend on the time.
At the risk of being tedious, but in the hope of avoiding future errors, let us
now treat the conguration A
0
= A
1
= A
2
= 0 and A
3
= Et. Since again

B = 0
and E
i
= E
i3
, this is the same physical problem, although at rst sight it looks
dierent, since now
H =
m

1 v
2
, (8.43)
but here H is not time-independent, since L depends on t explicitly, through A
3
, so

H = L/t ,= 0.
The Hamilton equations of motion are
m x
i

1 v
2
= p
i
+ eEt
i3
, (8.44)
together with

p
i
= 0, so now all the p
i
are time-independent, but v is not. Since
m

1 v
2
=

m
2
+
m
2
x
i
x
i
1 v
2
=
_
m
2
+ p
2
1
+ p
2
2
+ ( p
3
+ eEt)
2
, (8.45)
we can cast Eq. (8.44), i = 3, into the form
x
3
=
p
3
+ eEt
_
m
2
+ p
2
1
+ p
2
2
+ ( p
3
+ eEt)
2
. (8.46)
This equation can be integrated to give
(C + eEx
3
)
2
( p
3
+ eEt)
2
= C
2
p
2
3
, (8.47)
where C is an integration constant. This equation has the same algebraic form as
Eq. (8.39), the only dierence being that the constants have other names. We invite
the reader to complete the solution, and to show that the catenary, Eq. (8.41), can
be recovered.
It is instructive to observe that the two congurations are related by a gauge
transformation of the form Eq. (6.26). If

A = 0, and A
0
= Ex
3
(our rst example),
then it is sucient to take G = Ex
3
t, in Eq. (6.26), to see that A
0
= 0 and
A
3
=
3
G =
3
G = Et (our second example).
Although the above problem is relatively straightforward, two lessons can be
learned:
45
[1] The Hamiltonian method is easier if L/t = 0 ,
for then the Hamiltonian is a constant of the motion.
[2] One can sometimes use the gauge freedom to ensure that

H = 0.
46
Chapter 9
Radiation from a Moving Charge
9.1 Solution of the Wave Equation
In a region of space-time in which there are no sources (i.e. where j

= 0), the
Maxwell equations Eq. (6.29) in Lorentz gauge Eq. (6.27) reduce to

2
A

= 0 . (9.1)
Solutions of this equation can be found by Fourier transformation:
A

(x) =
_
d
3
ke
ikx
3

=0
a(

k, )

k, ), (9.2)
where kx = k

= k
0
t

k.x. The four polarization vectors, (

k, ), = 0, 1, 2, 3,
correspond to the four Lorentz components of A

. The spacelike components are


as in Appendix B, that is, they are mutually orthogonal unit spacelike vectors,
such that (

k, 3) is parallel to

k. (

k, 0) is a unit vector in the timelike direction,


(1, 0, 0, 0).
From Eq. (9.1) we have

2
A

=
_
d
3
ke
ikx
k
2
3

=0
a(

k, )

k, ) = 0 . (9.3)
Since the Fourier components are independent, k
2
= (k
0
)
2

k.

k = 0, or equivalently,
k
0
= [

k[ k . (9.4)
This constraint is called the mass-shell condition; it corresponds to the fact that
photons are massless.
The Lorentz condition Eq. (6.27) implies

=
_
d
3
ke
ikx
k
_
a(

k, 0) a(

k, 3)
_
= 0 ; (9.5)
and this means simply that a(

k, 0) = a(

k, 3). Under this constraint, any solution of


the Maxwell equations can be written in the form Eq. (9.2).
47
9.2 Retarded Potentials
The scalar potential engendered by a static charge, e, at a distance r is e/(4r).
If we have a static charge distribution, (x

), then the scalar potential at the point


x is given by the integral
(x ) =
1
4
_
d
3
x

(x

)
[x

x [
. (9.6)
If however the charge distribution is time-dependent, we expect the above integral
to be modied, since the potential should not be determined by the instantaneous
charge-distribution, but rather by the retarded charge-distribution, i.e. the distrib-
ution as it was when a light-signal was emitted from x

, that has just arrived at x .


Thus
(t, x ) =
1
4
_
d
3
x

(t [x

x [, x

)
[x

x [
. (9.7)
In Appendix E it is shown that Eq. (9.6) is indeed a solution of the dierential
equation

2
(x ) = (x ) ; (9.8)
while Eq. (9.7) is a solution of the time-dependent wave-equation

2
(t, x ) = (t, x ) , (9.9)
where
2
=

. The rather intuitive derivations given above are thus put on a


rm foundation. Moreover, since the Maxwell equation for the 4-potential has the
form

2
A

= j

, (9.10)
when the Lorentz gauge condition Eq. (6.27) is satised, it follows that we can use
the analysis of Appendix E to write the retarded solution to Eq. (9.10) in the form
A

(x) =
1
4
_
d
3
x

(t [x

x [, x

)
[x

x [
. (9.11)
This is not the most general solution of the Maxwell equation, Eq. (9.10), since any
solution of the homogeneous equation, Eq. (9.1), may be added to Eq. (9.11). The
retarded solution is assumed, however, to give the physical, or causal solution of the
Maxwell equation.
9.3 Lienard-Wiechert Fields
Let us specialize the general expression Eq. (9.11) for the retarded potential to
the case that the source is a single point charge, for example an electron, which is
at a position

X, that depends on the time. The corresponding charge density is
(x

) = j
0
(x

) = e
3
(x


X) , (9.12)
48
where e is the electrons charge and the -function is Diracs distribution, which
can be thought of as the limit of an innitely high, innitely sharply-peaked func-
tion, located at the point where its argument is zero. The corresponding current
distribution is
(x

) = ev
3
(x


X) , (9.13)
where v =

X, the velocity of the electron.


With these expressions for the charge and current densities, we can evaluate the
integral in Eq. (9.11). A diculty, however, is that the time retardation means that
the zeroth, or time component of the argument of j

depends on x

. The easiest
way to keep things straight is to introduce a t

-integration, and a time-retardation


delta-distribution. For the scalar potential, this yields
A
0
(x) =
e
4
_
d
4
x

3
(x


X (t

))(t

t +[x

x [)
1
[x

x [
=
e
4
_
dt

(t

t +[

X(t

) x [)
1
[

X(t

) x [
. (9.14)
Before the last integration is performed, the integration variable is changed from
t

to

t = t

t +[

X(t

) x [. However,
d

t =
_
1 +

t

X(t

) x [
_
dt

=
_
_
1 +
(

X(t

) x ).

X(t

)
[

X(t

) x [
_
_
dt

. (9.15)
In other words,
dt

= d

t/ , (9.16)
where
= 1 n.v , (9.17)
with
n =
x

X(t

)
[x

X(t

)[
. (9.18)
Thus n is a unit vector pointing from the point charge, as it was at the retarded
time, to the eld point, x . The integration over

t can now be performed trivially,
and we nd
A
0
(x) =
_
e
4R
_
ret
. (9.19)
Here
R = [x

X(t

)[ , (9.20)
49
and the sux ret means that t

, which occurs implicitly in the denition of and


of R, must be replaced by the retarded time, i.e. by the solution of the equation
t = t

+ R(t

) . (9.21)
Similarly, for the vector potential,

A(x) =
_
ev
4R
_
ret
. (9.22)
The expressions Eq. (9.19) and Eq. (9.22) are called the Lienard-Wiechert potentials.
They give the 4-potential at the point x , at time t, in terms of the position

X of
the point charge, as it was at the time t

= t [

X(t

) x [.
From the 4-potential, the Lienard-Wiechert elds can be calculated. They are

E =
e
4
3
_
n [(n v)

v]
R
+
(1 v
2
)(n v)
R
2
_
, (9.23)

B =
e
4
3
n
_

v +n (v

v)
R
+
(1 v
2
)v
R
2
_
. (9.24)
In these expressions, as in subsequent ones, the specication that the quantities on
the right are to be evaluated at the retarded time is implicit. The details of this
calculation are to be found in Appendix F, as is the proof that the above Lienard-
Wiechert elds are related by

B = n

E . (9.25)
As can be seen from the above expressions, both the electric eld and the mag-
netic induction have contributions that behave like R
2
, and others that behave like
R
1
. The former give the near-eld, and are all that is left if the acceleration of the
charge,

v , is zero:

E =
e(1 v
2
)
4
3
R
2
(n v) ; (9.26)
while

B can always be obtained from

E by means of Eq. (9.25). These terms
correspond to electrostatics and magnetostatics, i.e. to a charge moving at constant
speed and so giving rise to a constant current. The elds fall o according to the
famous inverse square law of distance. Indeed, with v << 1, and to lowest order
(O(1) for

E and O(v) for

B), and with

v = 0, we nd

E =
en
4R
2
, (9.27)

B =
en v
4R
2
. (9.28)
These terms are not surprising: they arise essentially from the application of a
Lorentz transformation to the static situation.
50
The really amazing thing about the Lienard-Wiechert elds lies in the R
1
or
radiation terms. These are present only if the charge undergoes acceleration. For
large R, we can neglect the R
2
terms, leaving

E =
e
4
3
R
n [(n v)

v] , (9.29)
together with Eq. (9.25). In the radiation zone (far from the accelerated source), the
electric eld is at right angles to the direction of the source (i.e. orthogonal to n),
and also to the magnetic induction, which is also at right angles to the direction of
the source. In a word, the radiation is transverse. That we have to do with radiation
is clear, since the energy and momentum densities are proportional to the squares
of the eld strengths, so to R
2
, and this means that there is a continual streaming
of energy and momentum into space. We shall illustrate this in the nonrelativistic
limit.
For nonrelativistic motion of the electron, i.e. v << 1, the Lienard-Wiechert
elds reduce to

E =
e
4R
n (n

v) +
e
4R
2
n , (9.30)

B =
e
4R
n

v . (9.31)
For large R, we neglect the static R
2
term, and we nd

E.

E =
_
en

v
4R
_
2
=

B.

B . (9.32)
Now according to Eq. (6.40), the energy density of the eld (the hamiltonian density)
is
H =
1
2
[

E.

E +

B.

B] +

E.

A
0
. (9.33)
The last term here does not contribute in the radiation zone. Indeed, even without
the nonrelativistic approximation, we can calculate from Eq. (9.19) that

E.

A
0
=
e(1 v
2
)
16
2

5
R
4
(n v).v . (9.34)
Hence the hamiltonian density in the radiation zone is
H =
e
2
v
2
sin
2

16
2
R
2
, (9.35)
where is the angle between n and

v, and this means that, in a shell about the
source of radius R and innitesimal thickness R, the instantaneous radiant energy
c is
c = 2
_
R+R
R
R
2
dR
_

0
sin d
e
2
v
2
sin
2

16
2
R
2
=
1
6
e
2
v
2
R . (9.36)
51
The energy in the shell is independent of R: it propagates outwards, without loss,
with the speed of light indeed it is light!! (Or radio waves, or IR, UV, -radiation,
or whatever!) Note that the R
2
terms in the elds do not contribute to this energy
in the limit of large R: they fall o too rapidly.
Consider lastly a collection of point charges, e
n
, located at

X
n
. We dene the
dipole moment of this collection of charges by

D =

n
e
n

X
n
. (9.37)
In the nonrelativistic limit, and in the radiation zone (i.e. v << 1 and R large com-
pared with the dimensions of the source), we have, from Eq. (9.30) and Eq. (9.31),

E =
n (n

D)
4R
, (9.38)
and

B =
n

D
4R
. (9.39)
As we have seen, the electric eld and the magnetic induction are transverse to the
propagation direction, are orthogonal to each other, and are equal in amplitude. At
an angle from the direction of

D, we have
E
2
= B
2
=

D
2
sin
2

16
2
R
2
. (9.40)
Thus the radiation is most intense at right angles to the second derivative of the
dipole moment, as anyone who has played with a radio transmission antenna knows!
52
Appendix A
Triple Vector Product
The vector product of two three-vectors can be written
(a

b )
i
=
ijk
a
j
b
k
, (A.1)
where the repeated indices, j and k, are summed over 1, 2, 3. This is a general rule:
if a Latin index is repeated in the same term, a summation is implied. The symbol,

ijk
, is dened to be +1 if i, j, k is a cyclic permutation of 1, 2, 3, and 1 for an
anticyclic permutation. It is 0 if two or more of the indices are the same. From
Eq. (A.1), we see that, for example, (a

b )
1
= a
2
b
3
a
3
b
2
.
A triple vector product can be calculated as follows:
[a (

b c )]
i
=
ijk
a
j

klm
b
l
c
m
=
ijk

lmk
a
j
b
l
c
m
= (
il

jm

im

jl
)a
j
b
l
c
m
= a
j
b
i
c
j
a
j
b
j
c
i
. (A.2)
The summation over j in the last line corresponds to a scalar product of two vectors.
In vector notation, we have shown that
a (

b c ) =

b (a .c ) c (a .

b ). (A.3)
From Eq. (A.2), we can immediately evaluate
[v (



A)]
i
= v
j

i
A
j
v
j

j
A
i
, (A.4)
Note that the order of the factors is important. Further,
[



A)]
i
=
i

j
A
j

j
A
i
. (A.5)
In vector notation, this last equation reads



A) =

(

A)
2

A. (A.6)
53
Appendix B
Potentials
B.1 Vector Potential
If a vector eld,

B(x), satises

.

B = 0, then there exists a vector eld,



A(x),
such that

B =


A.
Proof
Given

B(x), we write a three-dimensional Fourier transform as follows,

B(x) =
_
d
3
ke
ikx
3

=1
b(

k, ) (

k, ). (B.1)
Here (

k, ), = 1, 2, 3, are three orthonormal vectors, (

k, 3) being parallel to

k,
the others being therefore orthogonal to it. kx means

k.x, in this, and in subsequent


formulas. Clearly

B(x) = i
_
d
3
ke
ikx
kb(

k, 3), (B.2)
where k = [

k[. The necessary and sucient condition that



.

B = 0 is accordingly
that b(

k, 3) = 0, k ,= 0. Dene now

A(x) as follows:

A(x) =
_
d
3
ke
ikx
3

=1
a(

k, ) (

k, ), (B.3)
where a(

k, 1) = ib(

k, 2)/k, a(

k, 2) = ib(

k, 1)/k, for k ,= 0, while a(

k, 3) is arbitrary.
Then


A(x) = i
_
d
3
ke
ikx
3

=1
a(

k, )

k (

k, )
= i
_
d
3
ke
ikx
k[a(

k, 1) (

k, 2) a(

k, 2) (

k, 1)]
=
_
d
3
ke
ikx
[b(

k, 1) (

k, 1) + b(

k, 2) (

k, 2)]. (B.4)
The last expression can be seen to be equal to

B(x), since b(

k, 3) = 0, k ,= 0.
54
B.2 Scalar Potential
If a vector eld,

C(x), satises


C = 0, then there exists a scalar eld, (x),
such that

C =

.
Proof
In terms of the Fourier transform,

C(x) =
_
d
3
ke
ikx
3

=1
c(

k, ) (

k, ), (B.5)
we have


C(x) = i
_
d
3
ke
ikx
k[c(

k, 1) (

k, 2) c(

k, 2) (

k, 1)]. (B.6)
For this to vanish, it is necessary and sucient that c(

k, 1) = 0 = c(

k, 2), for k ,= 0.
Dene now the scalar eld, (x), by means of the Fourier transform
(x) =
_
d
3
ke
ikx
(

k), (B.7)
where (

k) = ic(

k, 3)/k, for k ,= 0. Then it follows that

(x) =
_
d
3
ke
ikx
c(

k, 3) (

k, 3). (B.8)
We recognize the right-hand side of the above equation as

C(x).
55
Appendix C
Invariant Measure
If x

and x are related by a Lorentz transformation, and F(x) is a Lorentz scalar,


then
_
d
4
x

F(x

) =
_
d
4
xF(x) (C.1)
Proof
By the usual rules for changing variables in a multiple integral, we have
_
d
4
x

F(x

) =
_
d
4
x
(x

0
, x

1
, x

2
, x

3
)
(x
0
, x
1
, x
2
, x
3
)
F(x) , (C.2)
where the Jacobian can be written
(x

0
, x

1
, x

2
, x

3
)
(x
0
, x
1
, x
2
, x
3
)
= det(

) , (C.3)
in which the matrix is dened by

=
x

, (C.4)
that is, it is the Lorentz transformation matrix itself. We shall show that the
determinant of this matrix is always unity, thus completing the proof of Eq. (C.1).
For a rotation about one of the space axes by an angle ,
det(

) = cos
2
+ sin
2
= 1 ; (C.5)
while a Lorentz transformation along one of the space axes by a hyperbolic boost,
u, gives
det(

) = cosh
2
u sinh
2
u = 1 . (C.6)
Since a general Lorentz transformation can be built up by compounding rotations
about the space axes (Euler angles), and a pure Lorentz boost along one space axis,
and since the determinant of the product of a number of matrices is equal to the
product of the determinants of each matrix, each of which is equal to unity, we see
nally that the determinant of the most general Lorentz transformation matrix is
also equal to unity.
56
Consider next an innitesimal Lorentz transformation along the x
1
axis, de-
scribed by a hyperbolic angle, u. The transformation matrix is

, (C.7)
where
0
1
= u =
1
0
, and the other elements of

are zero. For an innitesimal


rotation by an angle, , about the x
3
-axis, on the other hand, Eq. (C.7) holds, with

1
2
= =
2
1
, and the other elements are zero.
Now consider the doubly contravariant form,

= g

. (C.8)
In the case of the Lorentz transformation,

01
=
0

g
1
= u
10
=
1

g
0
= u , (C.9)
so that

is antisymmetric. For the rotation,

12
=
1

g
2
=
21
=
2

g
1
= , (C.10)
which means that here, too,

is antisymmetric.
Since any innitesimal Lorentz transformation can be made by compounding
innitesimal rotations around axes and a Lorentz transformation along an axis, it
follows that the general case can be written

= g

, (C.11)
where the innitesimal, , is antisymmetric, i.e.

. (C.12)
It is important to note that this antisymmetry applies to the doubly contravariant
form

, or of course equally well for the doubly covariant form

. It is not true
for the mixed form,

, in terms of which the Lorentz transformation was originally


dened.
57
Appendix D
Laplacian
Consider a transformation of Cartesian coordinates, x
1
, x
2
, x
3
, in three dimensions,
to another set of orthogonal coordinates, u
1
, u
2
, u
3
, such that
(dx
1
)
2
+ (dx
2
)
2
+ (dx
3
)
2
= (h
1
du
1
)
2
+ (h
2
du
2
)
2
+ (h
3
du
3
)
2
, (D.1)
where the hs are functions of the us. Let be an innitesimal volume. Then the
divergence of a vector,

A, can be written

A =
1

dx
1
dx
2
dx
3

.

A =
1

_
S
dS n.

A , (D.2)
where S is the surface of the innitesimal volume, and n is a unit vector normal to
this surface. In terms of the components in the u
1
, u
2
and u
3
directions, this can be
written
1

__
u
1
+du
1

_
u
1
_
h
2
du
2
h
3
du
3
A
1
+ permutations ,
where permutations means the two terms obtained by replacing the indices 1, 2, 3
by respectively 2, 3, 1 and 3, 1, 2 . This is, however,
1

_
u
1
du
1
du
2
du
3

u
1
[h
2
h
3
A
1
] + permutations ;
and since du
1
du
2
du
3
= (dx
1
dx
2
dx
3
)/(h
1
h
2
h
3
), it follows that

A =
1
h
1
h
2
h
3

u
1
[h
2
h
3
A
1
] + permutations . (D.3)
In the case that

A is the gradient of a scalar,

A =

, we nd

2
=
1
h
1
h
2
h
3
_

u
1
_
h
2
h
3
h
1

u
1
_
+

u
2
_
h
3
h
1
h
2

u
2
_
+

u
3
_
h
1
h
2
h
3

u
3
__
.(D.4)
The case of spherical polars is of particular interest. These are dened by
x
1
= r sin cos
x
2
= r sin sin
x
3
= r cos , (D.5)
58
which yields
(dx
1
)
2
+ (dx
2
)
2
+ (dx
3
)
2
= (dr)
2
+ (rd)
2
+ (r sin d)
2
, (D.6)
so that
h
1
= 1 , h
2
= r , h
3
= r sin . (D.7)
Hence we can write

2
=
1
r
2

r
r
2

r
+
1
r
2
sin

sin

+
1
r
2
sin
2

2
. (D.8)
A simple manipulation shows that the following operator identity holds:

2
=
1
r

2
r
2
r +
L
2
r
2
, (D.9)
where
L
2
=
1
sin

sin

+
1
sin
2

2
. (D.10)
59
Appendix E
Charge Distributions
E.1 Static Potential
Suppose that (x

) is a twice-dierentiable function, such that


_

0
r

dr

[(x

)[ < , (E.1)
where r

= [x

[. Dene
(x ) =
1
4
_
d
3
x

(x

)
[x

x [
. (E.2)
We shall show that

2
(x ) = (x ). (E.3)
E.1.1 Proof
Change the integration variable from x

to

R = x x

, so that
(x ) =
1
4
_
d
3
R
R
(x

R), (E.4)
where R = [

R[. Hence

2
(x ) =
1
4
_
d
3
R
R

2
(x

R), (E.5)
where
2
on the RHS can be shifted from the x to the

R variable. In terms of polar
coordinates,

2
=
1
R

2
R
2
R +
L
2
R
2
, (E.6)
where
L
2
=
1
sin

sin

+
1
sin
2

2
. (E.7)
60
The integral in Eq. (E.5) can be separated into three parts:

2
(x ) = (
1
+
2
+
3
)/(4), (E.8)
where

1
=
_
d
_

0
dR

2
R
2
[R(x

R)]
=
_
d[R

R
(x

R) + (x

R)]
R=0
= 4(x ), (E.9)

2
=
_

0
dR
R
_
2
0
d
_

0
d

sin

(x

R) = 0, (E.10)

3
=
_

0
dR
R
_

0
d
sin
_
2
0
d

2

2
(x

R) = 0. (E.11)
The result Eq. (E.3) follows on combining the last four equations.
There is a subtlety here that ought not to be glossed over. Suppose that we were
to apply the Laplacian to Eq. (E.2), without the change of integration variable:

2
(x ) =
1
4
_
d
3
x

(x

)
2
1
[x

x [
. (E.12)
Apparently nothing prevents us from rewriting the Laplacian in terms of R, and
, as in Eq. (E.6), but without making the change in integration variable. However,

2
1
[x

x [
=
1
R

2
R
2
R
1
R
= 0, (E.13)
so that one is sorely tempted to the erroneous conclusion that
2
vanishes.
Let us see what has gone wrong by considering the case that (x

) = 1 if [x

[ <
,and is zero otherwise. We can do an explicit calculation:
(x ) =
1
2
_

0
r
2
dr

_
1
1
d cos

r
2
+ r
2
2rr

cos
=
_

0
dr

r
2
max(r, r

)
=
1
2

1
6
r
2
, (E.14)
on condition that r < . For r > , we have clearly
(x ) =

3
3r
. (E.15)
It is now easy to check that
2
(x ) = 1 if r < . For r > , the Laplacian of
obviously vanishes.
The illegal step is Eq. (E.12), because if one performs the xdierentiations
under the integral, the resulting integral is not absolutely convergent, and the for-
mal vanishing of Eq. (E.13) does not guarantee the vanishing of the integral! The
objection clearly does not apply to Eq. (E.5).
61
We need to separate the integration domain into a sphere of radius around
[x

x [ = 0, and the rest. However, it is necessary to do this very smoothly. We


dene the test-function, f

R), by the requirements that it be unity for [

R[ < ,
and zero for [

R[ > 2. For < [

R[ < 2, we wish f

R) to go smoothly from 1 to 0.
If we require it to be innitely dierentiable in (0 < R < ), then all its derivatives
at R = and at at R = 2 must vanish.
Let us write

(x ) =
1
4
_
d
3
x

([x

x [)
(x

)
[x

x [
, (E.16)

(x ) =
1
4
_
d
3
x

[1 f

([x

x [)]
(x

)
[x

x [
. (E.17)
Clearly, from Eq. (E.2),
(x ) =

(x ) +

(x ). (E.18)
Moreover,

really is divergenceless, for Eq. (E.12) is unexceptionable, when a


sphere around the point x

= x is excised from the domain of the integral. Since


we have already demonstrated Eq. (E.3), by the legal proof Eq. (E.4) Eq. (E.11),
it follows that

(x ) = (x ), (E.19)
and this holds, no matter how small is.
This result is summarized by the formula

2
1
[x

x [
= 4
3
(x

x ), (E.20)
where
3
is the three-dimensional Dirac delta function (or more properly generalized
function, or distribution). This replaces Eq. (E.13). Within the more general frame-
work of distribution theory, Eq. (E.12) is allowed, with the identication Eq. (E.20).
E.2 Retarded Potential
If the charge-distribution is time-dependent, we write (x) (t, x ), and we
dene the retarded potential by
(t, x ) =
1
4
_
d
3
x

(t [x

x [, x

)
[x

x [
. (E.21)
We shall show that

2
(t, x ) = (t, x ). (E.22)
62
E.2.1 Proof
We rst dene

(t, x ) and

(t, x ) , analogously to Eq. (E.16) and Eq. (E.17),


i.e. in the former the integration domain of the integral in Eq. (E.21) is smoothly
restricted to a sphere of radius 2 about the point x

= x , while

(t, x ) contains
the rest of the integral. In the latter,

(t, x ) =
1
4
_
d
3
x

1
R

2
R
2
R[1 f

(R)]
(t R, x

)
R

=
1
4
_
d
3
x

[1 f

(R)]
1
R

2
t
2
(t R, x

) + O()
=

2

(t, x )
t
2
+ O(), (E.23)
i.e.
2

(t, x ) 0 as 0. As to the contribution of the small sphere about


R = 0, it is easy to see that
2

/t
2
tends to zero as 0. Moreover, just as in
the static case, we can apply the method of Eq. (E.8) to Eq. (E.11) to show that

= (t R, x

R)[
R=0
= (t, x ). (E.24)
This concludes the demonstration of Eq. (E.22).
In a completely analogous way, an advanced potential can be dened by
(t, x ) =
1
4
_
d
3
x

(t +[x

x [, x

)
[x

x [
, (E.25)
and it is clear that this, too, is a solution of Eq. (E.22). That the retarded, and
not the advanced solution must be used, in order to describe the radiation from an
accelerating charge, is a prescription that is not contained in the Maxwell dierential
equations: it is a deep question of boundary conditions.
63
Appendix F
Retarded Fields
We will sketch the rather tedious calculation of the Lienard-Wiechert elds, Eq. (9.23)
and Eq. (9.24), starting from the potentials, Eq. (9.19) and Eq. (9.22).
F.1 Lemmata
The following auxiliary theorems will rst be proved:
:
dt

dt
=
1

,
:

t

=
n

,
:
n
t

=
n(nv)
R
,
: [

(R)]
t

xed
= n v ,
:
(R)
t

= v
2
n.v Rn.

v .
Here t = t

+ R(t

) , R(t

) = [x

X(t

)[ , n =
x

X(t

)
R
.
Further = 1 n.v , and v =
d

X(t

)
dt

X .
F.1.1 Proofs
: dt = [1 +

R]dt

, and

R
R(t

)
t

=
x

X(t

)
R
.

X = n.v .
: 0 =

t =

t

+n.

(x

X) =

t

+n (n.v)

:

t

X
R
=

X
R

n
R

R =
(n.v)nv
R
:

(R) =

[R (x

X).v ] = n v
:
(R)
t

=

R
_
n
t

.v +n.

v
_
R = R(n.

v) (n v).v .
64
F.2 Theorems
The main theorems are:
I:

A
0
=
e

3
_
n.

v
R
+
1v
2
R
2
_
n +
e

2
R
2
v ,
II: [


A ]
t

xed
=
e

2
R
2
n v ,
III:
_


A
t

_
xxed
=
e

2
R
[

v +n (v

v )] +
e

2
R
2
[n.v v
2
]v .
F.2.1 Proofs
With the help of the Lemmata, and the following decompositions, the desired
formulae for the Lienard-Wiechert elds can be readily obtained:
I:

A
0
=
e

2
R
2
_
[

(R)]
t

xed
+
(R)
t

_
II: [


A]
t

xed
=
e

2
R
2

(R) v
III:
_


A
t

_
x xed
=
e

v
R

ev

2
R
2
(R)
t

In calculating

B from Theorem II, one must remember that
*


A = [


A ]
t

xed
+

t


A
t

_
x xed
.
Similarly, in using Theorem III to calculate

E, one uses
*


A
t
=
_


A
t

_
x xed
dt

dt
.
F.3 Corollary
From Eq. (9.23) we have
n

E =
e

3
R
n n [(n v)

v ] +
e(1v
2
)

3
R
2
n (n v)
However
n n

C = nn.

C

C.
So with

C = (n v)

v
we nd
nn[(nv)

v ] = nn.(v

v )n

v +v

v = n

v +n(v

v ).
Hence
n

E =
e

3
R
n

v +n (v

v )
e(1v
2
)

3
R
2
n v =

B.
As we see from Eq. (9.24).
65

Vous aimerez peut-être aussi