Electromagnetic Field Theory
Electromagnetic Field Theory
ϒ
Bo Thidé
U p s i l o n B o o k s
i i
i
i i
i
ϒ
E LECTROMAGNETIC F IELD T HEORY
Bo Thidé
i i
i
Also available
E LECTROMAGNETIC F IELD T HEORY
E XERCISES
by
Tobia Carozzi, Anders Eriksson, Bengt Lundborg,
Bo Thidé and Mattias Waldenvik
Freely downloadable from
www.plasma.uu.se/CED
i i
i
E LECTROMAGNETIC
F IELD T HEORY
Bo Thidé
ϒ
Upsilon Books · Communa AB · Uppsala · Sweden
i i
i
This book was typeset in LATEX 2ε (based on TEX 3.14159 and Web2C 7.4.2)
on an HP Visualize 9000/360 workstation running HP-UX 11.11.
i i
i
Contents
Preface xi
1 Classical Electrodynamics 1
1.1 Electrostatics . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.1.1 Coulomb’s law . . . . . . . . . . . . . . . . . . . . . . 2
1.1.2 The electrostatic field . . . . . . . . . . . . . . . . . . . 3
1.2 Magnetostatics . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.2.1 Ampère’s law . . . . . . . . . . . . . . . . . . . . . . . 6
1.2.2 The magnetostatic field . . . . . . . . . . . . . . . . . . 7
1.3 Electrodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.3.1 Equation of continuity for electric charge . . . . . . . . 10
1.3.2 Maxwell’s displacement current . . . . . . . . . . . . . 10
1.3.3 Electromotive force . . . . . . . . . . . . . . . . . . . . 11
1.3.4 Faraday’s law of induction . . . . . . . . . . . . . . . . 12
1.3.5 Maxwell’s microscopic equations . . . . . . . . . . . . 15
1.3.6 Maxwell’s macroscopic equations . . . . . . . . . . . . 16
1.4 Electromagnetic Duality . . . . . . . . . . . . . . . . . . . . . 16
Example 1.1 Faraday’s law as a consequence of conserva-
tion of magnetic charge . . . . . . . . . . . . . 18
Example 1.2 Duality of the electromagnetodynamic equations 19
Example 1.3 Dirac’s symmetrised Maxwell equations for a
fixed mixing angle . . . . . . . . . . . . . . . . 20
Example 1.4 The complex field six-vector . . . . . . . . . 21
Example 1.5 Duality expressed in the complex field six-vector 22
Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2 Electromagnetic Waves 25
2.1 The Wave Equations . . . . . . . . . . . . . . . . . . . . . . . 26
2.1.1 The wave equation for E . . . . . . . . . . . . . . . . . 26
2.1.2 The wave equation for B . . . . . . . . . . . . . . . . . 26
2.1.3 The time-independent wave equation for E . . . . . . . 27
Example 2.1 Wave equations in electromagnetodynamics . . 28
2.2 Plane Waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
2.2.1 Telegrapher’s equation . . . . . . . . . . . . . . . . . . 31
2.2.2 Waves in conductive media . . . . . . . . . . . . . . . . 32
i i
i
ii C ONTENTS
3 Electromagnetic Potentials 37
3.1 The Electrostatic Scalar Potential . . . . . .. . . . . . . . . . . 37
3.2 The Magnetostatic Vector Potential . . . . .. . . . . . . . . . . 38
3.3 The Electrodynamic Potentials . . . . . . .. . . . . . . . . . . 38
3.3.1 Lorenz-Lorentz gauge . . . . . . .. . . . . . . . . . . 40
The retarded potentials . . . . . . .. . . . . . . . . . . 44
3.3.2 Coulomb gauge . . . . . . . . . . .. . . . . . . . . . . 44
3.3.3 Gauge transformations . . . . . . .. . . . . . . . . . . 45
Example 3.1 Electromagnetodynamic potentials . . . . . . . 46
Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
4 Relativistic Electrodynamics 49
4.1 The Special Theory of Relativity . . . . . . . . . . . . . . . . . 49
4.1.1 The Lorentz transformation . . . . . . . . . . . . . . . 50
4.1.2 Lorentz space . . . . . . . . . . . . . . . . . . . . . . . 51
Radius four-vector in contravariant and covariant form . 52
Scalar product and norm . . . . . . . . . . . . . . . . . 52
Metric tensor . . . . . . . . . . . . . . . . . . . . . . . 53
Invariant line element and proper time . . . . . . . . . . 54
Four-vector fields . . . . . . . . . . . . . . . . . . . . . 56
The Lorentz transformation matrix . . . . . . . . . . . . 56
The Lorentz group . . . . . . . . . . . . . . . . . . . . 56
4.1.3 Minkowski space . . . . . . . . . . . . . . . . . . . . . 57
4.2 Covariant Classical Mechanics . . . . . . . . . . . . . . . . . . 59
4.3 Covariant Classical Electrodynamics . . . . . . . . . . . . . . . 61
4.3.1 The four-potential . . . . . . . . . . . . . . . . . . . . . 61
4.3.2 The Liénard-Wiechert potentials . . . . . . . . . . . . . 62
4.3.3 The electromagnetic field tensor . . . . . . . . . . . . . 64
Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
iii
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
iv C ONTENTS
F Formulae 163
F.1 The Electromagnetic Field . . . . . . . . . . . . . . . . . . . . 163
F.1.1 Maxwell’s equations . . . . . . . . . . . . . . . . . . . 163
Constitutive relations . . . . . . . . . . . . . . . . . . . 163
F.1.2 Fields and potentials . . . . . . . . . . . . . . . . . . . 163
Vector and scalar potentials . . . . . . . . . . . . . . . . 163
The Lorenz-Lorentz gauge condition in vacuum . . . . . 164
F.1.3 Force and energy . . . . . . . . . . . . . . . . . . . . . 164
Poynting’s vector . . . . . . . . . . . . . . . . . . . . . 164
Maxwell’s stress tensor . . . . . . . . . . . . . . . . . . 164
F.2 Electromagnetic Radiation . . . . . . . . . . . . . . . . . . . . 164
F.2.1 Relationship between the field vectors in a plane wave . 164
F.2.2 The far fields from an extended source distribution . . . 164
F.2.3 The far fields from an electric dipole . . . . . . . . . . . 164
F.2.4 The far fields from a magnetic dipole . . . . . . . . . . 165
F.2.5 The far fields from an electric quadrupole . . . . . . . . 165
F.2.6 The fields from a point charge in arbitrary motion . . . . 165
F.3 Special Relativity . . . . . . . . . . . . . . . . . . . . . . . . . 165
F.3.1 Metric tensor . . . . . . . . . . . . . . . . . . . . . . . 165
F.3.2 Covariant and contravariant four-vectors . . . . . . . . . 166
F.3.3 Lorentz transformation of a four-vector . . . . . . . . . 166
F.3.4 Invariant line element . . . . . . . . . . . . . . . . . . . 166
F.3.5 Four-velocity . . . . . . . . . . . . . . . . . . . . . . . 166
F.3.6 Four-momentum . . . . . . . . . . . . . . . . . . . . . 166
F.3.7 Four-current density . . . . . . . . . . . . . . . . . . . 166
F.3.8 Four-potential . . . . . . . . . . . . . . . . . . . . . . . 166
F.3.9 Field tensor . . . . . . . . . . . . . . . . . . . . . . . . 167
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
Appendices 163
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
vi C ONTENTS
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
List of Figures
vii
i i
i
i i
i
i i
i
i i
i
Preface
This book is the result of a twenty-five year long love affair. In 1972, I took
my first advanced course in electrodynamics at the Theoretical Physics depart-
ment, Uppsala University. Shortly thereafter, I joined the research group there
and took on the task of helping my supervisor, professor P ER -O LOF F RÖ -
MAN , with the preparation of a new version of his lecture notes on Electricity
Theory. These two things opened up my eyes for the beauty and intricacy of
electrodynamics, already at the classical level, and I fell in love with it.
Ever since that time, I have off and on had reason to return to electro-
dynamics, both in my studies, research and teaching, and the current book
is the result of my own teaching of a course in advanced electrodynamics at
Uppsala University some twenty odd years after I experienced the first en-
counter with this subject. The book is the outgrowth of the lecture notes that I
prepared for the four-credit course Electrodynamics that was introduced in the
Uppsala University curriculum in 1992, to become the five-credit course Clas-
sical Electrodynamics in 1997. To some extent, parts of these notes were based
on lecture notes prepared, in Swedish, by B ENGT L UNDBORG who created,
developed and taught the earlier, two-credit course Electromagnetic Radiation
at our faculty.
Intended primarily as a textbook for physics students at the advanced un-
dergraduate or beginning graduate level, I hope the book may be useful for
research workers too. It provides a thorough treatment of the theory of elec-
trodynamics, mainly from a classical field theoretical point of view, and in-
cludes such things as electrostatics and magnetostatics and their unification
into electrodynamics, the electromagnetic potentials, gauge transformations,
covariant formulation of classical electrodynamics, force, momentum and en-
ergy of the electromagnetic field, radiation and scattering phenomena, electro-
magnetic waves and their propagation in vacuum and in media, and covariant
Lagrangian/Hamiltonian field theoretical methods for electromagnetic fields,
particles and interactions. The aim has been to write a book that can serve
both as an advanced text in Classical Electrodynamics and as a preparation for
studies in Quantum Electrodynamics and related subjects.
In an attempt to encourage participation by other scientists and students in
the authoring of this book, and to ensure its quality and scope to make it useful
in higher university education anywhere in the world, it was produced within
a World-Wide Web (WWW) project. This turned out to be a rather successful
xi
i i
i
xii P REFACE
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
1
Classical
Electrodynamics
Classical electrodynamics deals with electric and magnetic fields and inter-
actions caused by macroscopic distributions of electric charges and currents.
This means that the concepts of localised electric charges and currents assume
the validity of certain mathematical limiting processes in which it is considered
possible for the charge and current distributions to be localised in infinitesim-
ally small volumes of space. Clearly, this is in contradiction to electromagnet-
ism on a truly microscopic scale, where charges and currents have to be treated
as spatially extended objects and quantum corrections must be included. How-
ever, the limiting processes used will yield results which are correct on small
as well as large macroscopic scales.
It took the genius of James Clerk Maxwell to unify electricity and magnet-
ism into a super-theory, electromagnetism or classical electrodynamics (CED),
and to realise that optics is a subfield of this new super-theory. Early in the 20th
century, Nobel laureate Hendrik Antoon Lorentz took the electrodynamics the-
ory further to the microscopic scale and also laid the foundation for the special
theory of relativity, formulated by Albert Einstein in 1905. In the 1930s Paul
A. M. Dirac expanded electrodynamics to a more symmetric form, includ-
ing magnetic as well as electric charges and also laid the foundation for the
development of quantum electrodynamics (QED) for which Sin-Itiro Tomon-
aga, Julian Schwinger, and Richard P. Feynman earned their Nobel prizes in
1965. Around the same time, physicists such as the Nobel laureates Sheldon
Glashow, Abdus Salam, and Steven Weinberg managed to unify electrodynam-
ics with the weak interaction theory to yet another super-theory, electroweak
theory.
In this chapter we start with the force interactions in classical electrostat-
ics and classical magnetostatics and introduce the static electric and magnetic
fields and find two uncoupled systems of equations for them. Then we see how
the conservation of electric charge and its relation to electric current leads to
the dynamic connection between electricity and magnetism and how the two
i i
i
2 C LASSICAL E LECTRODYNAMICS
1.1 Electrostatics
The theory which describes physical phenomena related to the interaction
between stationary electric charges or charge distributions in space with sta-
tionary boundaries is called electrostatics. For a long time electrostatics, un-
der the name electricity, was considered an independent physical theory of its
own, alongside other physical theories such as magnetism, mechanics, optics
and thermodynamics.1
where in the last step Formula (F.71) on page 169 was used. In SI units, which
we shall use throughout, the force F is measured in Newton (N), the electric
charges q and q in Coulomb (C) [= Ampère-seconds (As)], and the length
|x − x | in metres (m). The constant ε0 = 107 /(4πc2 ) ≈ 8.8542 × 10−12 Farad
1 The physicist and philosopher Pierre Duhem (1861–1916) once wrote:
‘The whole theory of electrostatics constitutes a group of abstract ideas and gen-
eral propositions, formulated in the clear and concise language of geometry and
algebra, and connected with one another by the rules of strict logic. This whole
fully satisfies the reason of a French physicist and his taste for clarity, simplicity
and order. . . .’
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
1.1 E LECTROSTATICS 3
x − x
x
q
x
per metre (F/m) is the vacuum permittivity and c ≈ 2.9979 × 10 8 m/s is the
speed of light in vacuum. In CGS units ε0 = 1/(4π) and the force is measured
in dyne, electric charge in statcoulomb, and length in centimetres (cm).
def F
Estat ≡ lim (1.2)
q→0 q
and Magnetism, first published in 1873, James Clerk Maxwell describes this in the following,
almost poetic, manner [9]:
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
4 C LASSICAL E LECTRODYNAMICS
Using (1.1) and Equation (1.2) on the previous page, and Formula (F.70)
on page 169, we find that the electrostatic field E stat at the field point x (also
known as the observation point), due to a field-producing electric charge q at
the source point x , is given by
q x − x q 1 q 1
E (x) =
stat
=− ∇ = ∇ (1.3)
4πε0 |x − x |3 4πε0 |x − x | 4πε0 |x − x |
In the presence of several field producing discrete electric charges q i , loc-
ated at the points xi , i = 1, 2, 3, . . . , respectively, in an otherwise empty space,
the assumption of linearity of vacuum 1 allows us to superimpose their indi-
vidual electrostatic fields into a total electrostatic field
1 x − xi
4πε0 ∑
Estat (x) = q (1.4)
i
i
x − x 3
i
If the discrete electric charges are small and numerous enough, we intro-
duce the electric charge density ρ, measured in C/m 3 in SI units, located at x
within a volume V of limited extent and replace summation with integration
over this volume. This allows us to describe the total field as
1 3 x−x 1 3 1
E (x) =
stat
d x ρ(x ) =− d x ρ(x )∇
4πε0 V |x − x |3 4πε0 V |x − x |
(1.5)
1 3 ρ(x )
=− ∇ dx
4πε0 V |x − x |
where we used Formula (F.70) on page 169 and the fact that ρ(x ) does not
depend on the unprimed (field point) coordinates on which ∇ operates.
We emphasise that under the assumption of linear superposition, Equa-
tion (1.5) is valid for an arbitrary distribution of electric charges, including
discrete charges, in which case ρ is expressed in terms of Dirac delta distribu-
tions:
ρ(x ) = ∑ qi δ(x − xi ) (1.6)
i
‘For instance, Faraday, in his mind’s eye, saw lines of force traversing all space
where the mathematicians saw centres of force attracting at a distance: Faraday
saw a medium where they saw nothing but distance: Faraday sought the seat of
the phenomena in real actions going on in the medium, they were satisfied that
they had found it in a power of action at a distance impressed on the electric
fluids.’
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
1.1 E LECTROSTATICS 5
x − xi
x
qi
V
xi
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
6 C LASSICAL E LECTRODYNAMICS
ρ(x)
∇ · Estat (x) = (1.9a)
ε0
∇ × E (x) = 0
stat
(1.9b)
1.2 Magnetostatics
While electrostatics deals with static electric charges, magnetostatics deals
with stationary electric currents, i.e., electric charges moving with constant
speeds, and the interaction between these currents. Here we shall discuss this
theory in some detail.
107 1
ε0 µ0 = 2
(F/m) × 4π × 10−7 (H/m) = 2 (s2 /m2 ) (1.11)
4πc c
which is a most useful relation.
At first glance, Equation (1.10) above may appear unsymmetric in terms
of the loops and therefore to be a force law which is in contradiction with
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
1.2 M AGNETOSTATICS 7
C dl
x − x dl
x
C
J
x
Newton’s third law. However, by applying the vector triple product ‘bac-cab’
Formula (F.51) on page 168, we can rewrite (1.10) as
µ0 JJ 1
F(x) = − dl dl · ∇
4π C C |x − x |
(1.12)
µ0 JJ x−x
− dl ·dl
4π C C |x − x |3
Since the integrand in the first integral is an exact differential, this integral van-
ishes and we can rewrite the force expression, Equation (1.10) on the preceding
page, in the following symmetric way
µ0 JJ x − x
F(x) = − dl · dl (1.13)
4π C C |x − x |3
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
8 C LASSICAL E LECTRODYNAMICS
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
1.3 E LECTRODYNAMICS 9
(1.18)
In the first of the two integrals on the right hand side, we use the representation
of the Dirac delta function given in Formula (F.73) on page 169, and integrate
the second one by parts, by utilising Formula (F.56) on page 168 as follows:
3 1
d x [j(x ) · ∇ ]∇
V |x − x |
3 ∂ 1 3
1
= x̂k d x ∇ · j(x ) − d x ∇ · j(x ) ∇
V ∂xk |x − x | V |x − x |
∂ 1
1
= x̂k dS · j(x )
− d3x ∇ · j(x ) ∇
S ∂xk |x − x | V |x − x |
(1.19)
Then we note that the first integral in the result, obtained by applying Gauss’s
theorem, vanishes when integrated over a large sphere far away from the loc-
alised source j(x ), and that the second integral vanishes because ∇ · j = 0 for
stationary currents (no charge accumulation in space). The net result is simply
∇×B stat
(x) = µ0 d3x j(x )δ(x − x ) = µ0 j(x) (1.20)
V
1.3 Electrodynamics
As we saw in the previous sections, the laws of electrostatics and magneto-
statics can be summarised in two pairs of time-independent, uncoupled vector
partial differential equations, namely the equations of classical electrostatics
ρ(x)
∇ · Estat (x) = (1.21a)
ε0
∇ × E (x) = 0
stat
(1.21b)
Since there is nothing a priori which connects E stat directly with Bstat , we must
consider classical electrostatics and classical magnetostatics as two independ-
ent theories.
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
10 C LASSICAL E LECTRODYNAMICS
2. A change in the magnetic flux through a loop will induce an EMF elec-
tric field in the loop. This is the celebrated Faraday’s law of induction.
∂ρ(t, x)
+ ∇ · j(t, x) = 0 (1.23)
∂t
which states that the time rate of change of electric charge ρ(t, x) is balanced
by a divergence in the electric current density j(t, x).
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
1.3 E LECTRODYNAMICS 11
where, in the last step, we have assumed that a generalisation of Equation (1.5)
on page 4 to time-varying fields allows us to make the identification
1 ∂ 3 1 ∂ 1 3 1
d x ρ(t, x )∇ = − d x ρ(t, x )∇
4πε0 ∂t V |x − x | ∂t 4πε0 V |x − x |
∂
= E(t, x)
∂t
(1.25)
Later, we will need to consider this formal result further. The result is Max-
well’s source equation for the B field
∂
∇ × B(t, x) = µ0 j(t, x) + ε0 E(t, x) (1.26)
∂t
where the last term ∂ε0 E(t, x)/∂t is the famous displacement current. This
term was introduced, in a stroke of genius, by Maxwell [8] in order to make
the right hand side of this equation divergence free when j(t, x) is assumed to
represent the density of the total electric current, which can be split up in ‘or-
dinary’ conduction currents, polarisation currents and magnetisation currents.
The displacement current is an extra term which behaves like a current density
flowing in vacuum. As we shall see later, its existence has far-reaching phys-
ical consequences as it predicts the existence of electromagnetic radiation that
can carry energy and momentum over very long distances, even in vacuum.
where σ is the electric conductivity (S/m). In the most general cases, for in-
stance in an anisotropic conductor, σ is a tensor.
We can view Ohm’s law, Equation (1.27) above, as the first term in a Taylor
expansion of the law j[E(t, x)]. This general law incorporates non-linear effects
such as frequency mixing. Examples of media which are highly non-linear are
semiconductors and plasma. We draw the attention to the fact that even in cases
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
12 C LASSICAL E LECTRODYNAMICS
when the linear relation between E and j is a good approximation, we still have
to use Ohm’s law with care. The conductivity σ is, in general, time-dependent
(temporal dispersive media) but then it is often the case that Equation (1.27)
on the preceding page is valid for each individual Fourier component of the
field.
If the current is caused by an applied electric field E(t, x), this electric field
will exert work on the charges in the medium and, unless the medium is super-
conducting, there will be some energy loss. The rate at which this energy is
expended is j · E per unit volume. If E is irrotational (conservative), j will
decay away with time. Stationary currents therefore require that an electric
field which corresponds to an electromotive force (EMF) is present. In the
presence of such a field EEMF , Ohm’s law, Equation (1.27) on the previous
page, takes the form
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
1.3 E LECTRODYNAMICS 13
dS
v
B(x)
dl
B(x)
where Φm is the magnetic flux and S is the surface encircled by C which can be
interpreted as a generic stationary ‘loop’ and not necessarily as a conducting
circuit. Application of Stokes’ theorem on this integral equation, transforms it
into the differential equation
∂
∇ × E(t, x) = −B(t, x) (1.32)
∂t
which is valid for arbitrary variations in the fields and constitutes the Maxwell
equation which explicitly connects electricity with magnetism.
Any change of the magnetic flux Φm will induce an EMF. Let us there-
fore consider the case, illustrated if Figure 1.3.4, that the ‘loop’ is moved in
such a way that it links a magnetic field which varies during the movement.
The convective derivative is evaluated according to the well-known operator
formula
d ∂
= +v·∇ (1.33)
dt ∂t
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
14 C LASSICAL E LECTRODYNAMICS
∇ · B(t, x) = 0 (1.35)
∇ × (B × v) = (v · ∇)B (1.36)
where EEMF is the field which is induced in the ‘loop’, i.e., in the moving
system. The use of Stokes’ theorem ‘backwards’ on Equation (1.39) above
yields
∂B
∇ × (EEMF − v × B) = − (1.40)
∂t
In the fixed system, an observer measures the electric field
E = EEMF − v × B (1.41)
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
1.3 E LECTRODYNAMICS 15
EEMF = E + (v × B) (1.43)
Hence, we can conclude that for a stationary observer, the Maxwell equation
∂B
∇×E = − (1.44)
∂t
is indeed valid even if the ‘loop’ is moving.
ρ(t, x)
∇·E = (1.45a)
ε0
∂B
∇×E = − (1.45b)
∂t
∇·B = 0 (1.45c)
∂E
∇ × B = ε0 µ0 + µ0 j(t, x) (1.45d)
∂t
In these equations ρ(t, x) represents the total, possibly both time and space de-
pendent, electric charge, i.e., free as well as induced (polarisation) charges,
and j(t, x) represents the total, possibly both time and space dependent, elec-
tric current, i.e., conduction currents (motion of free charges) as well as all
atomistic (polarisation, magnetisation) currents. As they stand, the equations
therefore incorporate the classical interaction between all electric charges and
currents in the system and are called Maxwell’s microscopic equations. An-
other name often used for them is the Maxwell-Lorentz equations. Together
with the appropriate constitutive relations, which relate ρ and j to the fields,
and the initial and boundary conditions pertinent to the physical situation at
hand, they form a system of well-posed partial differential equations which
completely determine E and B.
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
16 C LASSICAL E LECTRODYNAMICS
D = D[t, x; E, B] (1.46a)
H = H[t, x; E, B] (1.46b)
Under certain conditions, for instance for very low field strengths, we may
assume that the response of a substance to the fields is linear so that
D = εE (1.47)
H = µ−1 B (1.48)
i.e., that the derived fields are linearly proportional to the primary fields and
that the electric displacement (magnetising field) is only dependent on the elec-
tric (magnetic) field.
The field equations expressed in terms of the derived field quantities D and
H are
∇ · D = ρ(t, x) (1.49a)
∂B
∇×E = − (1.49b)
∂t
∇·B = 0 (1.49c)
∂D
∇×H = + j(t, x) (1.49d)
∂t
and are called Maxwell’s macroscopic equations. We will study them in more
detail in Chapter 6.
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
and make the ad hoc assumption that there exist magnetic monopoles repres-
ented by a magnetic charge density, which we denote by ρ m = ρm (t, x), and a
magnetic current density, which we denote by j m = jm (t, x). With these new
quantities included in the theory, and with the electric charge density denoted
ρe and the electric current density denoted j e , the Maxwell equations will be
symmetrised into the following four coupled, vector, partial differential equa-
tions:
ρe
∇·E = (1.50a)
ε0
∂B
∇×E = − − µ0 jm (1.50b)
∂t
1 ρm
∇ · B = µ0 ρm = 2 (1.50c)
c ε0
∂E 1 ∂E
∇ × B = ε0 µ0 + µ0 je = 2 + µ0 je (1.50d)
∂t c ∂t
We shall call these equations Dirac’s symmetrised Maxwell equations or the
electromagnetodynamic equations.
Taking the divergence of (1.50b), we find that
∂
∇ · (∇ × E) = − (∇ · B) − µ0 ∇ · jm ≡ 0 (1.51)
∂t
where we used the fact that, according to Formula (F.63) on page 168, the
divergence of a curl always vanishes. Using (1.50c) to rewrite this relation, we
obtain the equation of continuity for magnetic monopoles
∂ρm
+ ∇ · jm = 0 (1.52)
∂t
which has the same form as that for the electric monopoles (electric charges)
and currents, Equation (1.23) on page 10.
We notice that the new Equations (1.50) exhibit the following symmetry
(recall that ε0 µ0 = 1/c2 ):
E → cB (1.53a)
cB → −E (1.53b)
cρ → ρ
e m
(1.53c)
ρ → −cρ
m e
(1.53d)
cj → j
e m
(1.53e)
j → −cj
m e
(1.53f)
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
18 C LASSICAL E LECTRODYNAMICS
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
We assume that Formula (1.56) on the preceding page is valid also for time-varying
magnetic currents. Then, with the use of the representation of the Dirac delta function,
Equation (F.73) on page 169, the equation of continuity for magnetic charge, Equa-
tion (1.52) on page 17, and the assumption of the generalisation of Equation (1.55) on
the facing page to time-dependent magnetic charge distributions, we obtain, formally,
3 m µ0 ∂ 3 m 1
∇ × E(t, x) = −µ0 d x j (t, x )δ(x − x ) − d x ρ (t, x )∇
V 4π ∂t V |x − x |
∂
= −µ0 jm (t, x) − B(t, x)
∂t
(1.59)
[cf. Equation (1.24) on page 10] which we recognise as Equation (1.50b) on page 17.
A transformation of this electromagnetodynamic result by rotating into the ‘electric
realm’ of charge space, thereby letting jm tend to zero, yields the electrodynamic
Equation (1.50b) on page 17, i.e., the Faraday law in the ordinary Maxwell equations.
This process also provides an alternative interpretation of the term ∂B∂t as a magnetic
displacement current, dual to the electric displacement current [cf. Equation (1.26) on
page 11].
By postulating the indestructibility of a hypothetical magnetic charge, we have thereby
been able to replace Faraday’s experimental results on electromotive forces and induc-
tion in loops as a foundation for the Maxwell equations by a more appealing one.
E ND OF EXAMPLE 1.1
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
20 C LASSICAL E LECTRODYNAMICS
ρe
∇ · E = ∇ · (E cos θ + cB sin θ) = cos θ + cµ0 ρm sin θ
ε
0 e (1.60)
1 1 ρ
= ρe cos θ + ρm sin θ =
ε0 c ε0
∂ B ∂ 1
∇ × E + = ∇ × (E cos θ + cB sin θ) + − E sin θ + B cos θ
∂t ∂t c
∂B 1 ∂E
= −µ0 jm cos θ − cos θ + cµ0 je sin θ + sin θ
∂t c ∂t (1.61)
1 ∂E ∂B
− sin θ + cos θ = −µ0 jm cos θ + cµ0 je sin θ
c ∂t ∂t
= −µ0 (−cje sin θ + jm cos θ) = −µ0 jm
1 ρe
∇ · B = ∇ · (− E sin θ + B cos θ) = − sin θ + µ0 ρm cos θ
c cε0 (1.62)
= µ0 −cρe sin θ + ρm cos θ = µ0 ρm
1 ∂ E 1 1 ∂
∇ × B − = ∇ × (− E sin θ + B cos θ) − 2 (E cos θ + cB sin θ)
c2 ∂t c c ∂t
1 1 ∂B 1 ∂E
= µ0 jm sin θ + cos θ + µ0 je cos θ + 2 cos θ
c c ∂t c ∂t
(1.63)
1 ∂E 1 ∂B
− 2 cos θ − sin θ
c ∂t c ∂t
1 m
= µ0 j sin θ + je cos θ = µ0 je
c
QED
E ND OF EXAMPLE 1.2
E XAMPLE 1.3 D IRAC ’ S SYMMETRISED M AXWELL EQUATIONS FOR A FIXED MIXING ANGLE
Show that for a fixed mixing angle θ such that
ρm = cρe tan θ (1.64a)
jm = cje tan θ (1.64b)
the symmetrised Maxwell equations reduce to the usual Maxwell equations.
Explicit application of the fixed mixing angle conditions on the duality transformation
(1.54) on page 18 yields
e 1 1
ρ = ρe cos θ + ρm sin θ = ρe cos θ + cρe tan θ sin θ
c c (1.65a)
1 1 e
= (ρ cos θ + ρ sin θ) =
e 2 e 2
ρ
cos θ cos θ
m
ρ = −cρe sin θ + cρe tan θ cos θ = −cρe sin θ + cρe sin θ = 0 (1.65b)
e 1 e 1 e
j = je cos θ + je tan θ sin θ = (j cos2 θ + je sin2 θ) = j (1.65c)
cos θ cos θ
m
j = −cje sin θ + cje tan θ cos θ = −cje sin θ + cje sin θ = 0 (1.65d)
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
Hence, a fixed mixing angle, or, equivalently, a fixed ratio between the electric and
magnetic charges/currents, ‘hides’ the magnetic monopole influence (ρm and jm ) on
the dynamic equations.
We notice that the inverse of the transformation given by Equation (1.54) on page 18
yields
Furthermore, from the expressions for the transformed charges and currents above, we
find that
ρe 1 ρe
∇ · E = = (1.68)
ε0 cos θ ε0
and
∇ · B = µ0 ρm = 0 (1.69)
so that
1 ρe ρe
∇·E = cos θ − 0 = (1.70)
cos θ ε0 ε0
E ND OF EXAMPLE 1.3
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
22 C LASSICAL E LECTRODYNAMICS
E ND OF EXAMPLE 1.4
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
1.4 B IBLIOGRAPHY 23
Furthermore, assuming that θ = θ(t, x), we see that the spatial and temporal differenti-
ation of G leads to
∂ G
∂t G ≡ = −i(∂t θ)e−iθ G + e−iθ ∂t G (1.80a)
∂t
∂ · G ≡ ∇ · G = −ie−iθ ∇θ · G + e−iθ ∇ · G (1.80b)
−iθ −iθ
∂ × G ≡ ∇ × G = −ie ∇θ × G + e ∇×G (1.80c)
which means that ∂t G transforms as G itself only if θ is time-independent, and that
∇ · G and ∇ × G transform as G itself only if θ is space-independent.
E ND OF EXAMPLE 1.5
Bibliography
[1] T. W. BARRETT AND D. M. G RIMES, Advanced Electromagnetism. Found-
ations, Theory and Applications, World Scientific Publishing Co., Singapore,
1995, ISBN 981-02-2095-2.
[4] E. H ALLÉN, Electromagnetic Theory, Chapman & Hall, Ltd., London, 1962.
[5] J. D. JACKSON, Classical Electrodynamics, third ed., John Wiley & Sons, Inc.,
New York, NY . . . , 1999, ISBN 0-471-30932-X.
[6] L. D. L ANDAU AND E. M. L IFSHITZ, The Classical Theory of Fields, fourth re-
vised English ed., vol. 2 of Course of Theoretical Physics, Pergamon Press, Ltd.,
Oxford . . . , 1975, ISBN 0-08-025072-6.
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
24 C LASSICAL E LECTRODYNAMICS
[7] F. E. L OW, Classical Field Theory, John Wiley & Sons, Inc., New York, NY . . . ,
1997, ISBN 0-471-59551-9.
[14] J. S CHWINGER, A magnetic model of matter, Science, 165 (1969), pp. 757–761.
[17] J. VANDERLINDE, Classical Electromagnetic Theory, John Wiley & Sons, Inc.,
New York, Chichester, Brisbane, Toronto, and Singapore, 1993, ISBN 0-471-
57269-1.
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
2
Electromagnetic
Waves
ρ(t, x)
∇·E = (Coulomb’s/Gauss’s law) (2.1a)
ε0
∂B
∇×E = − (Faraday’s law) (2.1b)
∂t
∇·B = 0 (No free magnetic charges) (2.1c)
∂E
∇ × B = ε0 µ0 + µ0 j(t, x) (Ampère’s/Maxwell’s law) (2.1d)
∂t
25
i i
i
26 E LECTROMAGNETIC WAVES
∇ × (∇ × E) = ∇(∇ · E) − ∇2 E (2.3)
∇·E = 0 (2.4)
and since EEMF = 0, Ohm’s law, Equation (1.28) on page 12, yields
j = σE (2.5)
∂E 1 ∂2 E
∇2 E − µ0 σ − =0 (2.7)
∂t c2 ∂t2
which is the homogeneous wave equation for E.
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
which, with the use of Equation (F.64) on page 168 and Equation (2.1c) on
page 25 can be rewritten
∂B ∂2
∇(∇ · B) − ∇2 B = −µ0 σ − ε0 µ0 2 B (2.9)
∂t ∂t
Using the fact that, according to (2.1c), ∇ · B = 0 for any medium and rearran-
ging, we can rewrite this equation as
∂B 1 ∂2 B
∇2 B − µ0 σ − =0 (2.10)
∂t c2 ∂t2
This is the wave equation for the magnetic field. We notice that it is of exactly
the same form as the wave equation for the electric field, Equation (2.7) on the
facing page.
c
ω 2 σ
= ∇2 E + 2 1 + i E=0
c ε0 ω
Introducing the relaxation time τ = ε 0 /σ of the medium in question we can
rewrite this equation as
ω2 i
∇ E+ 2 1+
2
E=0 (2.13)
c τω
In the limit of long τ, Equation (2.13) tends to
ω2
∇2 E + E=0 (2.14)
c2
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
28 E LECTROMAGNETIC WAVES
∇2 E + iωµ0 σE = 0 (2.15)
ρe
∇·E = (2.19a)
ε0
∂B
∇×E = − − µ0 jm (2.19b)
∂t
∇ · B = µ0 ρm (2.19c)
∂E
∇ × B = ε 0 µ0 + µ0 je (2.19d)
∂t
under the assumption of vanishing net electric and magnetic charge densities and in
the absence of electromotive and magnetomotive forces. Interpret this equation phys-
ically.
Taking the curl of (2.19b) and using (2.19d), and assuming, for symmetry reasons, that
there exists a linear relation between the magnetic current density jm and the magnetic
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
30 E LECTROMAGNETIC WAVES
This, together with (2.27a), shows that the longitudinal component of E, i.e.,
the component which is perpendicular to the plane surface is independent of ζ
and has a time dependence which exhibits an exponential decay, with a decre-
ment given by the relaxation time τ in the medium.
Scalar multiplying (2.27b) by n̂, we similarly find that
∂E ∂B
0 = n̂· n̂× = − n̂· (2.31)
∂ζ ∂t
or
∂B
n̂· =0 (2.32)
∂t
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
From this, and (2.27c), we conclude that the only longitudinal component of
B must be constant in both time and space. In other words, the only non-static
solution must consist of transverse components.
∂2 E ∂E 1 ∂2 E
− µ0 σ − =0 (2.33)
∂ζ 2 ∂t c2 ∂t2
∂2 E 1 ∂2 E
− =0 (2.34)
∂ζ 2 c2 ∂t2
As is well known, each component of this equation has a solution which can
be written
E = E0 ei(k·x−ωt) (2.38)
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
32 E LECTROMAGNETIC WAVES
β2 = α2 − k2 (2.45)
k2 σ
αβ = (2.46)
2ε0 ω
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
Squaring the latter and combining with the former, one obtains the second
order algebraic equation (in α 2 )
k4 σ2
α2 (α2 − k2 ) = (2.47)
4ε20 ω2
which can be easily solved and one finds that
2
1+ σ +1
ε0 ω
α=k (2.48a)
2
2
1+ σ −1
ε0 ω
β=k (2.48b)
2
As a consequence, the solution of the time-independent telegrapher’s equation,
Equation (2.41) on the preceding page, can be written
E = E0 e−βζ ei(αζ−ωt) (2.49)
With the aid of Equation (2.40) on the facing page we can calculate the asso-
ciated magnetic field, and find that it is given by
1 1 1
B= K k̂ × E = ( k̂ × E)(α + iβ) = ( k̂ × E) |A| eiγ (2.50)
ω ω ω
where we have, in the last step, rewritten α + iβ in the amplitude-phase form
|A| exp{iγ}. From the above, we immediately see that E, and consequently also
B, is damped, and that E and B in the wave are out of phase.
In the case that ε0 ω σ, we can approximate K as follows:
1 1
σ 2 σ ε0 ω 2 σ
K = k 1+i =k i 1−i ≈ k(1 + i)
ε0 ω ε0 ω σ 2ε0 ω
(2.51)
√ σ µ0 σω
= ε0 µ0 ω(1 + i) = (1 + i)
2ε0 ω 2
From this analysis we conclude that when the wave impinges perpendicu-
larly upon the medium, the fields are given, inside this medium, by
µ0 σω µ0 σω
E = E0 exp − ζ exp i ζ − ωt (2.52a)
2 2
µ0 σ
B = (1 + i) ( n̂× E ) (2.52b)
2ω
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
34 E LECTROMAGNETIC WAVES
Furthermore, letting ∗ denotes complex conjugate, we can express the real part
of the complex vector F as
1
Re {F} = Re F0 (x) e−iωt = [F0 (x) e−iωt + F∗0 (x) eiωt ] (2.55)
2
and similarly for G. Hence, the physically acceptable interpretation of the
scalar product of two complex vectors, representing physical observables, is
F(t, x) · G(t, x) = Re F0 (x) e−iωt · Re G0 (x) e−iωt
1 1
= [F0 (x) e−iωt + F∗0 (x) eiωt ] · [G0 (x) e−iωt + G∗0 (x) eiωt ]
2 2
1
= F0 · G∗0 + F∗0 · G0 + F0 · G0 e−2iωt + F∗0 · G∗0 e2iωt
4
1 (2.56)
= Re F0 · G∗0 + F0 · G0 e−2iωt
2
1
= Re F0 e−iωt · G∗0 eiωt + F0 · G0 e−2iωt
2
1
= Re F(t, x) · G∗ (t, x) + F0 · G0 e−2iωt
2
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
2.3 B IBLIOGRAPHY 35
Bibliography
[1] J. D. JACKSON, Classical Electrodynamics, third ed., John Wiley & Sons, Inc.,
New York, NY . . . , 1999, ISBN 0-471-30932-X.
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
36 E LECTROMAGNETIC WAVES
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
3
Electromagnetic
Potentials
Taking the divergence of this and using Equation (1.7) on page 5, we obtain
Poisson’s equation
ρ(x)
∇2 φstat (x) = −∇ · Estat (x) = − (3.2)
ε0
where the integration is taken over all source points x at which the charge
density ρ(x ) is non-zero and α is an arbitrary quantity which has a vanish-
ing gradient. An example of such a quantity is a scalar constant. The scalar
function φstat (x) in Equation (3.3) is called the electrostatic scalar potential.
37
i i
i
38 E LECTROMAGNETIC P OTENTIALS
where a(x) is an arbitrary vector field whose curl vanishes. From Equa-
tion (F.62) on page 168 we know that such a vector can always be written
as the gradient of a scalar field.
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
Inserting this expression into the other homogeneous Maxwell equation, Equa-
tion (1.32) on page 13, we obtain
∂ ∂
∇ × E(t, x) = − [∇ × A(t, x)] = −∇ × A(t, x) (3.7)
∂t ∂t
or, rearranging the terms,
∂
∇ × E(t, x) + A(t, x) = 0 (3.8)
∂t
∂
E(t, x) + A(t, x) = −∇φ(t, x) (3.9)
∂t
This means that in electrodynamics, E(t, x) can be calculated from the formula
∂
E(t, x) = −∇φ(t, x) − A(t, x) (3.10)
∂t
and B(t, x) from Equation (3.6) on the preceding page. Hence, it is a matter
of taste whether we want to express the laws of electrodynamics in terms of
the potentials φ(t, x) and A(t, x), or in terms of the fields E(t, x) and B(t, x).
However, there exists an important difference between the two approaches: in
classical electrodynamics the only directly observable quantities are the fields
themselves (and quantities derived from them) and not the potentials. On the
other hand, the treatment becomes significantly simpler if we use the potentials
in our calculations and then, at the final stage, use Equation (3.6) on the facing
page and Equation (3.10) above to calculate the fields or physical quantities
expressed in the fields.
Inserting (3.10) and (3.6) on the facing page into Maxwell’s equa-
tions (1.45) on page 15 we obtain, after some simple algebra and the use of
Equation (1.11) on page 6, the general inhomogeneous wave equations
ρ(t, x) ∂
∇2 φ = − − (∇ · A) (3.11a)
ε0 ∂t
1∂ A
2 1 ∂φ
∇2 A − 2 2 − ∇(∇ · A) = −µ0 j(t, x) + 2 ∇ (3.11b)
c ∂t c ∂t
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
40 E LECTROMAGNETIC P OTENTIALS
1 ∂φ
∇·A+ =0 (3.13)
c2 ∂t
the coupled inhomegeneous wave Equation (3.12) on page 40 simplify into the
following set of uncoupled inhomogeneous wave equations:
def 1 ∂2 1 ∂2 φ ρ(t, x)
φ ≡
2
− ∇ 2
φ = − ∇2 φ = (3.14a)
c2 ∂t2 c2 ∂t2 ε0
def 1 ∂2 1 ∂2 A
2 A ≡ − ∇2 A = 2 2 − ∇2 A = µ0 j(t, x) (3.14b)
c ∂t
2 2 c ∂t
1 Infact, the Dutch physicist Hendrik Antoon Lorentz, who in 1903 demonstrated the cov-
ariance of Maxwell’s equations, was not the original discoverer of this condition. It had been
discovered by the Danish physicist Ludvig V. Lorenz already in 1867 [5]. In the literature, this
fact has sometimes been overlooked and the condition was earlier referred to as the Lorentz
gauge condition.
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
exists, and that the same is true for the generic potential component:
∞
Ψ(t, x) = dω Ψω (x) e−iωt (3.17a)
−∞
1 ∞
Ψω (x) = dt Ψ(t, x) eiωt (3.17b)
2π −∞
Inserting the Fourier representations (3.16a) and (3.17a) into Equation (3.15),
and using the vacuum dispersion relation for electromagnetic waves
ω = ck (3.18)
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
42 E LECTROMAGNETIC P OTENTIALS
and the solution of Equation (3.19) on the preceding page which corresponds
to the frequency ω is given by the superposition
Ψω (x) = d3x fω (x )G(x, x ) (3.21)
V
d2
(rG) + k2 (rG) = −rδ(r) (3.22)
dr2
Away from r = |x − x | = 0, i.e., away from the source point x , this equation
takes the form
d2
(rG) + k2 (rG) = 0 (3.23)
dr2
with the well-known general solution
eikr e−ikr eik|x−x | −e
−ik|x−x |
G = C+ + C− ≡ C+ + C (3.24)
r r |x − x | |x − x |
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
In virtue of the fact that the volume element d 3x in spherical polar coordinates
is proportional to |x − x |2 , the second integral vanishes when |x − x | → 0. Fur-
thermore, from Equation (F.73) on page 169, we find that the integrand in the
first integral can be written as −4πδ(|x − x |) and, hence, that
1
C+ + C− = (3.27)
4π
Insertion of the general solution Equation (3.24) on the facing page into
Equation (3.21) on the preceding page gives
+ 3 eik|x−x | e−ik|x−x |
Ψω (x) = C d x fω (x ) + C− 3
d x fω (x ) (3.28)
V |x − x | V |x − x |
The Fourier transform to ordinary t domain of this is obtained by inserting the
above expression for Ψω (x) into Equation (3.17a) on page 41:
|
∞ exp −iω t − k|x−x ω
Ψ(t, x) = C + d3x dω fω (x )
V −∞ |x − x |
|
(3.29)
∞ exp −iω t + k|x−x ω
+ C − d3x dω fω (x )
V −∞ |x − x |
and the advanced time t in the following
If we introduce the retarded time t ret adv
way [using the fact that in vacuum k/ω = 1/c, according to Equation (3.18) on
page 41]:
k |x − x | |x − x |
tret
= tret (t, x − x ) = t − = t− (3.30a)
ω c
k |x − x
| |x − x |
tadv = tadv (t, x − x ) = t +
= t+ (3.30b)
ω c
and use Equation (3.16a) on page 41, we obtain
, x ) , x )
+ f (tret
3 f (tadv
Ψ(t, x) = C dx + C− d3x (3.31)
V |x − x | V |x − x |
This is a solution to the generic inhomogeneous wave equation for the potential
components Equation (3.15) on page 41. We note that the solution at time t at
the field point x is dependent on the behaviour at other times t of the source
at x and that both retarded and advanced t are mathematically acceptable
solutions. However, if we assume that causality requires that the potential at
, x ), we must in
(t, x) is set up by the source at an earlier time, i.e., at (t ret
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
44 E LECTROMAGNETIC P OTENTIALS
1 ∂2 A 1 ∂φ
∇2 A − = −µ0 j(t, x) + 2 ∇ (3.34)
c2 ∂t2 c ∂t
The first of these two is the time-dependent Poisson’s equation which, in ana-
logy with Equation (3.3) on page 37, has the solution
1 ρ(t, x )
φ(t, x) = d3x +α (3.35)
4πε0 V |x − x |
where α has vanishing gradient. We note that in the scalar potential expression
the charge density source is evaluated at time t. The retardation (and advance-
ment) effects ccur only in the vector potential, which is the solution of the
1 In fact, inspired by a discussion by Paul A. M. Dirac, John A. Wheeler and Richard P. Feyn-
man derived in 1945 a fully self-consistent electrodynamics using both the retarded and the
advanced potentials [7]; see also [3].
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
where, once again Equation (F.62) on page 168 was used. Comparing these
expressions with (3.10) and (3.6) we see that the fields are unaffected by the
gauge transformation (3.37). A transformation of the potentials φ and A which
leaves the fields, and hence Maxwell’s equations, invariant is called a gauge
transformation. A physical law which does not change under a gauge trans-
formation is said to be gauge invariant. By definition, the fields themselves
are, of course, gauge invariant.
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
46 E LECTROMAGNETIC P OTENTIALS
The potentials φ(t, x) and A(t, x) calculated from (3.11a) on page 39, with
an arbitrary choice of ∇ · A, can be further gauge transformed according to
(3.37) on the previous page. If, in particular, we choose ∇ · A according
to the Lorenz-Lorentz condition, Equation (3.13) on page 40, and apply the
gauge transformation (3.37) on the resulting Lorenz-Lorentz potential equa-
tions (3.14) on page 40, these equations will be transformed into
1 ∂2 φ ∂ 1 ∂2 Γ ρ(t, x)
−∇ φ+
2
−∇ Γ =
2
(3.39a)
c2 ∂t2 ∂t c2 ∂t2 ε0
1 ∂2 A 1 ∂2 Γ
− ∇ 2
A − ∇ − ∇2 Γ = µ0 j(t, x) (3.39b)
c2 ∂t2 c2 ∂t2
We notice that if we require that the gauge function Γ(t, x) itself be restricted
to fulfil the wave equation
1 ∂2 Γ
− ∇2 Γ = 0 (3.40)
c2 ∂t2
these transformed Lorenz-Lorentz equations will keep their original form. The
set of potentials which have been gauge transformed according to Equation (3.37)
on the preceding page with a gauge function Γ(t, x) which is restricted to fulfil
Equation (3.40) above, i.e., those gauge transformed potentials for which the
Lorenz-Lorentz equations (3.14) are invariant, comprises the Lorenz-Lorentz
gauge.
The process of choosing a particular gauge condition is referred to as gauge
fixing.
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
∂ e
E = −∇φe (t, x) − A (t, x) − ∇ × Am (3.42a)
∂t
1 1 ∂
B = − 2 ∇φm (t, x) − 2 Am (t, x) + ∇ × Ae (3.42b)
c c ∂t
In the absence of magnetic charges, or, equivalenty for φm ≡ 0 and Am ≡ 0, these
formulae reduce to the usual Maxwell theory Formula (3.10) on page 39 and For-
mula (3.6) on page 38, respectively, as they should.
Inserting the symmetrised expressions (3.42) on the preceding page into Equa-
tions (3.41) on the facing page, one obtains [cf., Equations (3.11a) on page 39]
∂
ρe (t, x)
∇2 φe + ∇ · Ae = − (3.43a)
∂t ε0
∂
ρm (t, x)
∇2 φm + ∇ · Am = − (3.43b)
∂t ε0
1 ∂2 Ae 1 ∂φe
−∇ A +∇ ∇·A + 2
2 e e
= µ0 je (t, x) (3.43c)
c2 ∂t2 c ∂t
1 ∂2 Am 1 ∂φm
−∇ A +∇ ∇·A + 2
2 m m
= µ0 jm (t, x) (3.43d)
c2 ∂t2 c ∂t
1 ∂ e
∇ · Ae + φ =0 (3.44)
c2 ∂t
1 ∂
∇ · Am + 2 φm = 0 (3.45)
c ∂t
these coupled wave equations simplify to
1 ∂2 φe ρe (t, x)
− ∇ φ
2 e
= (3.46a)
c2 ∂t2 ε0
1 ∂2 φm ρm (t, x)
− ∇2 φm = (3.46b)
c2 ∂t2 ε0
1 ∂ A
2 e
− ∇2 Ae = µ0 je (t, x) (3.46c)
c2 ∂t2
1 ∂2 Am
− ∇2 Am = µ0 jm (t, x) (3.46d)
c2 ∂t2
exhibiting once again, the striking properties of Dirac’s symmetrised Maxwell theory.
E ND OF EXAMPLE 3.1
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
48 E LECTROMAGNETIC P OTENTIALS
Bibliography
[1] L. D. FADEEV AND A. A. S LAVNOV, Gauge Fields: Introduction to Quantum
Theory, No. 50 in Frontiers in Physics: A Lecture Note and Reprint Series. Ben-
jamin/Cummings Publishing Company, Inc., Reading, MA . . . , 1980, ISBN 0-
8053-9016-2.
[4] J. D. JACKSON, Classical Electrodynamics, third ed., John Wiley & Sons, Inc.,
New York, NY . . . , 1999, ISBN 0-471-30932-X.
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
4
Relativistic
Electrodynamics
49
i i
i
50 R ELATIVISTIC E LECTRODYNAMICS
vt
y y
Σ Σ v
P(t, x, y, z)
P(t , x , y , z )
O x O x
z z
F IGURE 4.1: Two inertial systems Σ and Σ in relative motion with velo-
city v along the x = x axis. At time t = t = 0 the origin O of Σ coincided
with the origin O of Σ. At time t, the inertial system Σ has been translated
a distance vt along the x axis in Σ. An event represented by P(t, x, y, z) in
Σ is represented by P(t , x , y , z ) in Σ .
Postulate 4.1 (Relativity principle; Poincaré, 1905). All laws of physics (ex-
cept the laws of gravitation) are independent of the uniform translational mo-
tion of the system on which they operate.
Postulate 4.2 (Einstein, 1905). The velocity of light in empty space is inde-
pendent of the motion of the source that emits the light.
A consequence of the first postulate is that all geometrical objects (vec-
tors, tensors) in an equation describing a physical process must transform in a
covariant manner, i.e., in the same way.
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
where v = |v|. In the following, we shall make frequent use of these shorthand
notations.
As shown by Einstein, the two postulates of special relativity require that
the spatial coordinates and times as measured by an observer in Σ and Σ ,
respectively, are connected by the following transformation:
Taking the difference between the square of (4.3a) and the square of (4.3b) we
find that
c2 t2 − x2 = γ2 c2 t2 − 2xcβt + x2 β2 − x2 + 2xvt − v2 t2
1 v2 v2
= c t 1− 2 − x 1− 2
2 2 2
v2 c c (4.4)
1− 2
c
= c2 t 2 − x 2
From Equations (4.3) we see that the y and z coordinates are unaffected by
the translational motion of the inertial system Σ along the x axis of system
Σ. Using this fact, we find that we can generalise the result in Equation (4.4)
above to
which means that if a light wave is transmitted from the coinciding origins O
and O at time t = t = 0 it will arrive at an observer at (x, y, z) at time t in Σ
and an observer at (x , y , z ) at time t in Σ in such a way that both observers
conclude that the speed (spatial distance divided by time) of light in vacuum
is c. Hence, the speed of light in Σ and Σ is the same. A linear coordin-
ate transformation which has this property is called a (homogeneous) Lorentz
transformation.
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
52 R ELATIVISTIC E LECTRODYNAMICS
speed of light and t is time), and the remaining components are the components
of the ordinary R3 radius vector x defined in Equation (M.1) on page 172:
xµ = gµν xν (4.7)
gµν xν xµ = xµ xµ (4.8)
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
the norm in a 4D Riemannian space, then the explicit expression for the scalar
product of xµ with itself in this space must be
xµ x µ = c2 t 2 − x 2 − y2 − z2 (4.9)
We notice that our space will have an indefinite norm which means that we deal
with a non-Euclidean space. We call the four-dimensional space (or space-
time) with this property Lorentz space and denote it L 4 . A corresponding real,
linear 4D space with a positive definite norm which is conserved during ordin-
ary rotations is a Euclidean vector space. We denote such a space R 4 .
Metric tensor
By choosing the metric tensor in L4 as
⎧
⎪
⎨1 if µ = ν = 0
gµν = −1 if µ = ν = i = j = 1, 2, 3 (4.10)
⎪
⎩
0 if µ = ν
i.e., a matrix with a main diagonal that has the sign sequence, or signature,
{+, −, −, −}, the index lowering operation in our chosen flat 4D space becomes
nearly trivial:
Hence, if the metric tensor is defined according to expression (4.10) above the
covariant radius four-vector x µ is obtained from the contravariant radius four-
vector xµ simply by changing the sign of the last three components. These
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
54 R ELATIVISTIC E LECTRODYNAMICS
ds2 = gµν dxν dxµ = dxµ dxµ = (dx0 )2 − (dx1 )2 − (dx2 )2 − (dx3 )2 (4.17)
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
where we introduced
dτ = dt/γ (4.19)
Since dτ measures the time when no spatial changes are present, it is called the
proper time.
Expressing Equation (4.5) on page 51 in terms of the differential interval
ds and comparing with Equation (4.17) on the facing page, we find that
is a light-like interval; we may also say that in this case we are on the light
cone. A vector which has a light-like interval is called a null vector. The
time-like, space-like or light-like aspects of an interval ds is invariant under a
Lorentz transformation.
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
56 R ELATIVISTIC E LECTRODYNAMICS
Four-vector fields
Any quantity which relative to any coordinate system has a quadruple of real
numbers and transforms in the same way as the radius four-vector x µ does, is
called a four-vector. In analogy with the notation for the radius four-vector we
introduce the notation aµ = (a0 , a) for a general contravariant four-vector field
in L4 and find that the ‘lowering of index’ rule, Equation (M.32) on page 178,
for such an arbitrary four-vector yields the dual covariant four-vector field
The scalar product between this four-vector field and another one b µ (xκ ) is
which is a scalar field, i.e., an invariant scalar quantity α(x κ ) which depends
on time and space, as described by x κ = (ct, x, y, z).
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
X0
X 0
θ
x1
θ
x1
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
58 R ELATIVISTIC E LECTRODYNAMICS
Σ w
x0 = ct
x0 x0 = x1
P
x1
ϕ ct
O= O
P x1 = x
√
where i = −1, we see that Equation (4.17) on page 54 transforms into
dS 2 = (dX 0 )2 + (dX 1 )2 + (dX 2 )2 + (dX 3 )2 (4.31)
i.e., into a 4D differential form which is positive definite just as is ordinary 3D
Euclidean space R3 . We shall call the 4D Euclidean space constructed in this
way the Minkowski space M4 .1
As before, it suffices to consider the simplified case where the relative
motion between Σ and Σ is along the x axes. Then
dS 2 = (dX 0 )2 + (dx1 )2 (4.32)
and we consider X 0 and x1 as orthogonal axes in an Euclidean space. As in all
Euclidean spaces, every interval is invariant under a rotation of the X 0 x1 plane
through an angle θ into X 0 x1 :
X 0 = −x1 sin θ + X 0 cos θ (4.33a)
1
x = x cos θ + X sin θ
1 0
(4.33b)
1 Thefact that our Riemannian space can be transformed in this way into an Euclidean one
means that it is, strictly speaking, a pseudo-Riemannian space.
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
which, when multiplied with the scalar invariant m 0 yields the four-momentum
⎛ ⎞
µ
dx m0 c m0 v ⎠
pµ = m0 = m0 γ(c, v) = ⎝ ( ,( = (p0 , p) (4.38)
dτ 1− v 2
1− v2
c2 c2
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
60 R ELATIVISTIC E LECTRODYNAMICS
p = mv (4.39)
where
m0
m = γm0 = ( (4.40)
2
1 − vc2
We can interpret this such that the Lorentz covariance implies that the mass-
like term in the ordinary 3D linear momentum is not invariant. A better way to
look at this is that p = mv = γm0 v is the covariantly correct expression for the
kinetic three-momentum.
Multiplying the zeroth (time) component of the four-momentum p µ with
the scalar invariant c, we obtain
m0 c2
cp0 = γm0 c2 = ( = mc2 (4.41)
2
1 − vc2
Since this component has the dimension of energy and is the result of a covari-
ant description of the motion of a particle with its kinetic momentum described
by the spatial components of the four-momentum, Equation (4.38) on the pre-
vious page, we interpret cp0 as the total energy E. Hence,
Since this is an invariant, this equation holds in any inertial frame, particularly
in the frame where p = 0 and there we have
E = m0 c2 (4.44)
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
ρ = γρ0 (4.46)
where φ is the scalar potential and A the vector potential, defined in Sec-
tion 3.3 on page 38, we can write the uncoupled inhomogeneous wave equa-
tions, Equations (3.14) on page 40, in the following compact (and covariant)
way:
2 Aµ = µ0 jµ (4.49)
With the help of the above, we can formulate our electrodynamic equations
covariantly. For instance, the covariant form of the equation of continuity,
Equation (1.23) on page 10 is
∂µ jµ = 0 (4.50)
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
62 R ELATIVISTIC E LECTRODYNAMICS
and the Lorenz-Lorentz gauge condition, Equation (3.13) on page 40, can be
written
∂µ Aµ = 0 (4.51)
If only one dimension Lorentz contracts (for instance, due to relative mo-
tion along the x direction), a 3D spatial volume transforms according to
1 v2
dV = d x = dV0 = dV0 1 − β = dV0 1 − 2
3 2 (4.53)
γ c
then from Equation (4.46) on the preceding page we see that
i.e., the charge in a given volume is conserved. We can therefore conclude that
the elementary charge is a universal constant.
where |x − x |0 is the usual distance from the source point to the field point,
evaluated in the rest system (signified by the index ‘0’).
Let us introduce the relative radius four-vector between the source point
and the field point:
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
We know that in vacuum the signal (field) from the charge q at xµ propag-
ates to xµ with the speed of light c so that
x − x = c(t − t ) (4.58)
Inserting this into Equation (4.57) on the facing page, we see that
Rµ Rµ = 0 (4.59)
Now we want to find the correspondence to the rest system solution, Equa-
tion (4.55) on the facing page, in an arbitrary inertial system. We note from
Equation (4.37) on page 59 that in the rest system
⎛ ⎞
µ c v ⎠
u 0 = ⎝( ,( = (c, 0) (4.61)
2 2
1 − vc2 1 − vc2
v=0
and
(Rµ )0 = (x − x , x − x )0 = (x − x 0 , (x − x )0 ) (4.62)
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
64 R ELATIVISTIC E LECTRODYNAMICS
uν Rν = γcs (4.68)
and
uµ 1 v
= , (4.69)
cuν Rν cs c2 s
from which we see that the solution (4.64) can be written
µ κ q 1 v φ
A (x ) = , = ,A (4.70)
4πε0 cs c2 s c
where in the last step the definition of the four-potential, Equation (4.48) on
page 61, was used. Writing the solution in the ordinary 3D-way, we conclude
that for a very localised charge volume, moving relative an observer with a
velocity v, the scalar and vector potentials are given by the expressions
q 1 q 1
φ(t, x) = = (4.71a)
4πε0 s 4πε0 |x − x | − β · (x − x )
q v q v
A(t, x) = = (4.71b)
4πε0 c s 4πε0 c |x − x | − β · (x − x )
2 2
ck = ai b j − a j bi = ci j = −c ji , i, j = k (4.73)
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
B = ∇×A (4.76a)
∂A
E = −∇φ − (4.76b)
∂t
In component form, this can be written
∂A j ∂Ai
Bi j = − = ∂i A j − ∂ j Ai (4.77a)
∂xi ∂x j
∂φ ∂Ai
Ei = − i − = −∂i φ − ∂t Ai (4.77b)
∂x ∂t
From this, we notice the clear difference between the axial vector (pseudovector)
B and the polar vector (‘ordinary vector’) E.
Our goal is to express the electric and magnetic fields in a tensor form
where the components are functions of the covariant form of the four-potential,
Equation (4.48) on page 61:
µ φ
A = ,A (4.78)
c
Inspection of (4.78) and Equation (4.77) makes it natural to define the four-
tensor
∂Aν ∂Aµ
F µν = − = ∂µ Aν − ∂ν Aµ (4.79)
∂xµ ∂xν
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
66 R ELATIVISTIC E LECTRODYNAMICS
can be written
⎛ ⎞
0 −E x /c −Ey /c −Ez /c
µν ⎜E x /c 0 −Bz By ⎟
F =⎜
⎝ Ey /c
⎟ (4.80)
Bz 0 −B x ⎠
Ez /c −By Bx 0
The covariant field tensor is obtained from the contravariant field tensor in the
usual manner by index contraction (index lowering):
∂ν F νµ = µ0 jµ (4.83)
∂F 01 ∂F 11 ∂F 21 ∂F 31 1 ∂E x ∂Bz ∂By
+ 1 + 2 + 3 =− 2 +0− + = µ0 j1 = µ0 ρv x (4.86)
∂x 0 ∂x ∂x ∂x c ∂t ∂y ∂z
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
4.3 B IBLIOGRAPHY 67
Hence, Equation (4.83) on the preceding page and Equation (4.91) constitute
Maxwell’s equations in four-dimensional formalism.
Bibliography
[1] J. A HARONI, The Special Theory of Relativity, second, revised ed., Dover Pub-
lications, Inc., New York, 1985, ISBN 0-486-64870-2.
[4] D. B OHM, The Special Theory of Relativity, Routledge, New York, NY, 1996,
ISBN 0-415-14809-X.
[6] L. D. L ANDAU AND E. M. L IFSHITZ, The Classical Theory of Fields, fourth re-
vised English ed., vol. 2 of Course of Theoretical Physics, Pergamon Press, Ltd.,
Oxford . . . , 1975, ISBN 0-08-025072-6.
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
68 R ELATIVISTIC E LECTRODYNAMICS
[7] F. E. L OW, Classical Field Theory, John Wiley & Sons, Inc., New York, NY . . . ,
1997, ISBN 0-471-59551-9.
[8] H. M UIRHEAD, The Special Theory of Relativity, The Macmillan Press Ltd.,
London, Beccles and Colchester, 1973, ISBN 333-12845-1.
[9] C. M ØLLER, The Theory of Relativity, second ed., Oxford University Press,
Glasgow . . . , 1972.
[12] B. S PAIN, Tensor Calculus, third ed., Oliver and Boyd, Ltd., Edinburgh and
London, 1965, ISBN 05-001331-9.
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
5
Electromagnetic
Fields and
Particles
69
i i
i
Lagrange formalism
Let us now introduce a generalised action
S (4) = L(4) (xµ , uµ ) dτ (5.1)
where dτ is the proper time defined via Equation (4.18) on page 55, and L (4)
acts as a kind of generalisation to the common 3D Lagrangian so that the vari-
ational principle
τ1
δS (4) (τ0 , τ1 ) = δ L(4) (xµ , uµ ) dτ = 0 (5.2)
τ0
1
free
L(4) = m0 uµ uµ (5.3)
2
For an interaction with the electromagnetic field we can introduce the interac-
tion with the help of the four-potential given by Equation (4.78) on page 65 in
the following way
1
L(4) = m0 uµ uµ + quµ Aµ (xν ) (5.4)
2
We call this the four-Lagrangian and shall now show how this function, to-
gether with the variation principle, Formula (5.2), yields covariant results which
are physically correct.
The variation principle (5.2) with the 4D Lagrangian (5.4) inserted, leads
to
τ1
m0 µ
δS (4) (τ0 , τ1 ) = δ u uµ + quµ Aµ dτ
τ0 2
τ1 µ
m0 ∂(u uµ ) µ µ µ ∂Aµ ν
= δu + q Aµ δu + u δx dτ (5.5)
τ0 2 ∂uµ ∂xν
τ1
= m0 uµ δuµ + q Aµ δuµ + uµ ∂ν Aµ δxν dτ = 0
τ0
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
in the integrand in the right hand member of Equation (5.12) on the preceding
page must vanish. In other words, we have found an equation of motion for a
charged particle in a prescribed electromagnetic field:
duν
m0 = quµ ∂ν Aµ − ∂µ Aν (5.13)
dτ
With the help of Equation (4.79) on page 65 we can express this equation in
terms of the electromagnetic field tensor in the following way:
duν
m0 = quµ Fνµ (5.14)
dτ
This is the sought-for covariant equation of motion for a particle in an electro-
magnetic field. It is often referred to as the Minkowski equation. As the reader
can easily verify, the spatial part of this 4-vector equation is the covariant (re-
lativistically correct) expression for the Newton-Lorentz force equation.
Hamiltonian formalism
The usual Hamilton equations for a 3D space are given by Equation (M.105)
on page 190 in Appendix M. These six first-order partial differential equations
are
∂H dqi
= (5.15a)
∂pi dt
∂H dpi
=− (5.15b)
∂qi dt
∂L(4)
pµ = (5.16)
∂uµ
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
With the help of these, the radius four-vector x µ , considered as the generalised
four-coordinate, and the invariant line element ds, defined in Equation (4.18)
on page 55, we introduce the following eight partial differential equations:
∂H(4) dxµ
= (5.18a)
∂pµ dτ
∂H(4) dpµ
=− (5.18b)
∂xµ dτ
which form the four-dimensional Hamilton equations.
Our strategy now is to use Equation (5.16) on the preceding page and Equa-
tions (5.18) to derive an explicit algebraic expression for the canonically con-
jugate momentum four-vector. According to Equation (4.42) on page 60, c
times a four-momentum has a zeroth (time) component which we can identify
with the total energy. Hence we require that the component p 0 of the conjug-
ate four-momentum vector defined according to Equation (5.16) on the facing
page be identical to the ordinary 3D Hamiltonian H divided by c and hence
that this cp0 solves the Hamilton equations, Equations (5.15) on the preceding
page. This later consistency check is left as an exercise to the reader.
Using the definition of H(4) , Equation (5.17) on the facing page, and the
expression for L(4) , Equation (5.4) on page 70, we obtain
1
H(4) = pµ uµ − L(4) = pµ uµ − m0 uµ uµ − quµ Aµ (xν ) (5.19)
2
Furthermore, from the definition (5.16) of the canonically conjugate four-
momentum pµ , we see that
µ ∂L(4) ∂ 1
p = = m0 u uµ + quµ A (x ) = m0 uµ + qAµ
µ µ ν
(5.20)
∂uµ ∂uµ 2
Inserting this into (5.19), we obtain
1 1
H(4) = m0 uµ uµ + qAµ uµ − m0 uµ uµ − quµ Aµ (xν ) = m0 uµ uµ (5.21)
2 2
Since the four-velocity scalar-multiplied by itself is u µ uµ = c2 , we clearly
see from Equation (5.21) that H(4) is indeed a scalar invariant, whose value is
simply
m0 c2
H(4) = (5.22)
2
However, at the same time (5.20) provides the algebraic relationship
1
µ
uµ = p − qAµ (5.23)
m0
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
ηi−1 ηi ηi+1
m m m m m
k k k k
a a a a x
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
1 N
2
L= ∑
2 i=1
mη̇i − k(ηi+1 − ηi )2 (5.37)
Here,
η − η 2
1 m 2 i+1 i
Li = η̇i − ka (5.39)
2 a a
is the so called linear Lagrange density. If we now let N → ∞ and, at the same
time, let the springs become infinitesimally short according to the following
scheme:
a → dx (5.40a)
m dm
→ =µ linear mass density (5.40b)
a dx
ka → Y Young’s modulus (5.40c)
ηi+1 − ηi ∂η
→ (5.40d)
a ∂x
we obtain
L= L dx (5.41)
where
& 2 '
2
∂η ∂η 1 ∂η ∂η
L η, , , t = µ −Y (5.42)
∂t ∂x 2 ∂t ∂x
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
Notice how we made a transition from a discrete description, in which the mass
points were identified by a discrete integer variable i = 1, 2, . . . , N, to a continu-
ous description, where the infinitesimal mass points were instead identified by
a continuous real parameter x, namely their position along x̂.
A consequence of this transition is that the number of degrees of freedom
for the system went from the finite number N to infinity! Another consequence
is that L has now become dependent also on the partial derivative with respect
to x of the ‘field coordinate’ η. But, as we shall see, the transition is well
worth the price because it allows us to treat all fields, be it classical scalar
or vectorial fields, or wave functions, spinors and other fields that appear in
quantum physics, on an equal footing.
Under the assumption of time independence and fixed endpoints, the vari-
ation principle (5.36) on the preceding page yields:
δ L dt
∂η ∂η
=δ L η, ,
dx dt
∂t ∂x
⎡ ⎤ (5.43)
⎣ ∂L ∂L ∂η ∂L ∂η ⎦ dx dt
= δη + δ + δ
∂η ∂ ∂η ∂t ∂ ∂η ∂x
∂t ∂x
=0
The last integral can be integrated by parts. This results in the expression
⎡ ⎛ ⎞ ⎛ ⎞⎤
⎣ ∂L − ∂ ⎝ ∂L ∂ ∂L
⎠ − ⎝ ⎠⎦ δη dx dt = 0 (5.44)
∂η ∂t ∂ ∂η ∂x ∂ ∂η
∂t ∂x
where the variation is arbitrary (and the endpoints fixed). This means that the
integrand itself must vanish. If we introduce the functional derivative
⎛ ⎞
δL ∂L ∂ ∂L
= − ⎝ ⎠ (5.45)
δη ∂η ∂x ∂ ∂η
∂x
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
i.e., the one-dimensional wave equation for compression waves which propag-
√
ate with phase speed vφ = Y/µ along the linear structure.
A generalisation of the above 1D results to a three-dimensional continuum
is straightforward. For this 3D case we get the variational principle
δ L dt = δ L d3x dt
∂η
= δ L η, µ d4x
∂x
⎡ ⎛ ⎞⎤ (5.48)
= ⎣ ∂L − ∂ ⎝ ∂L ⎠⎦ δη d4x
∂η ∂xµ ∂ ∂ηµ
∂x
=0
where the variation δη is arbitrary and the endpoints are fixed. This means that
the integrand itself must vanish:
⎛ ⎞
∂L ∂ ⎝ ∂L ⎠
− =0 (5.49)
∂η ∂xµ ∂ ∂ηµ
∂x
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
∂L
π(xµ ) = π(t, x) = (5.52)
∂ ∂η
∂t
If, as usual, we differentiate this expression and identify terms, we obtain the
following Hamilton density equations
∂H ∂η
= (5.54a)
∂π ∂t
δH ∂π
=− (5.54b)
δη ∂t
The Hamilton density functions are in many ways similar to the ordinary
Hamilton functions and lead to similar results.
where the mechanical part has to do with the particle motion (kinetic energy).
It is given by L(4) /V where L(4) is given by Equation (5.3) on page 70 and V is
the volume. Expressed in the rest mass density 0 , the mechanical Lagrange
density can be written
1
L mech = 0 uµ uµ (5.56)
2
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
The L inter part which describes the interaction between the charged particles
and the external electromagnetic field. A convenient expression for this inter-
action Lagrange density is
L inter = jµ Aµ (5.57)
For the field part L field we choose the difference between magnetic and
electric energy density (in analogy with the difference between kinetic and
potential energy in a mechanical field). Using the field tensor, we express this
field Lagrange density as
1 µν
L field = F Fµν (5.58)
4µ0
1 1 µν
L tot = 0 uµ uµ + jµ Aµ + F Fµν (5.59)
2 4µ0
where µ denotes the row number and ν the column number. Then, Einstein summation
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
or
1 µν 1 B2 1 2 1 B2
F Fµν = − E = − ε0 E 2 (5.64)
4µ0 2 µ0 c 2 µ0 2 µ0
where, in the last step, the identity ε0 µ0 = 1/c2 was used. QED
E ND OF EXAMPLE 5.1
1 µν
L EM = L inter + L field = jν Aν + F Fµν (5.65)
4µ0
inserted into the Euler-Lagrange equations, expression (5.49) on page 79, yields
two of Maxwell’s equations. To see this, we note from Equation (5.65) and the
results in Example 5.1 that
∂L EM
= jν (5.66)
∂Aν
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
Furthermore,
∂L EM 1 ∂
κλ
∂µ = ∂µ F Fκλ
∂(∂µ Aν ) 4µ0 ∂(∂µ Aν )
1 ∂
κ λ λ κ
= ∂µ (∂ A − ∂ A )(∂κ Aλ − ∂λ Aκ )
4µ0 ∂(∂µ Aν )
1 ∂
= ∂µ ∂κ Aλ ∂κ Aλ − ∂κ Aλ ∂λ Aκ (5.67)
4µ0 ∂(∂µ Aν )
λ κ λ κ
− ∂ A ∂κ Aλ + ∂ A ∂λ Aκ
1 ∂
κ λ κ λ
= ∂µ ∂ A ∂κ Aλ − ∂ A ∂λ Aκ
2µ0 ∂(∂µ Aν )
But
∂
κ λ ∂ ∂
∂ A ∂κ Aλ = ∂κ Aλ ∂κ Aλ + ∂κ Aλ ∂κ Aλ
∂(∂µ Aν ) ∂(∂µ Aν ) ∂(∂µ Aν )
∂ ∂
= ∂κ Aλ ∂κ Aλ + ∂κ Aλ gκα ∂α gλβ Aβ
∂(∂µ Aν ) ∂(∂µ Aν )
∂ ∂
= ∂κ Aλ ∂κ Aλ + gκα gλβ ∂κ Aλ ∂α Aβ (5.68)
∂(∂µ Aν ) ∂(∂µ Aν )
∂ ∂
= ∂κ Aλ ∂κ Aλ + ∂α Aβ ∂α Aβ
∂(∂µ Aν ) ∂(∂µ Aν )
= 2∂µ Aν
Similarly,
∂
κ λ
∂ A ∂λ Aκ = 2∂ν Aµ (5.69)
∂(∂µ Aν )
so that
∂L EM 1
1
∂µ = ∂µ ∂µ Aν − ∂ν Aµ = ∂µ F µν (5.70)
∂(∂µ Aν ) µ0 µ0
This means that the Euler-Lagrange equations, expression (5.49) on page 79,
for the Lagrangian density L EM and with Aν as the field quantity become
∂L EM ∂L EM 1
− ∂µ = jν − ∂µ F µν = 0 (5.71)
∂Aν ∂(∂µ Aν ) µ0
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
or
∂µ F µν = µ0 jν (5.72)
Other fields
In general, the dynamic equations for most any fields, and not only electro-
magnetic ones, can be derived from a Lagrangian density together with a vari-
ational principle (the Euler-Lagrange equations). Both linear and non-linear
fields are studied with this technique. As a simple example, consider a real,
scalar field η which has the following Lagrange density:
1
L = ∂µ η∂µ η − m2 η2 (5.73)
2
Insertion into the 1D Euler-Lagrange equation, Equation (5.46) on page 78,
yields the dynamic equation
(2 − m2 )η = 0 (5.74)
e−m|x|
η = ei(k·x−ωt) (5.75)
|x|
which describes the Yukawa meson field for a scalar meson with mass m. With
1 ∂η
π= (5.76)
c2 ∂t
we obtain the Hamilton density
1
2 2
H = c π + (∇η)2 + m2 η2 (5.77)
2
which is positive definite.
Another Lagrangian density which has attracted quite some interest is the
Proca Lagrangian
1 µν
L EM = L inter + L field = jν Aν + F Fµν + m2 Aµ Aµ (5.78)
4µ0
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
5.2 B IBLIOGRAPHY 85
∂µ F µν + m2 Aν = µ0 jν (5.79)
Bibliography
[1] A. O. BARUT, Electrodynamics and Classical Theory of Fields and Particles,
Dover Publications, Inc., New York, NY, 1980, ISBN 0-486-64038-8.
[5] L. D. L ANDAU AND E. M. L IFSHITZ, The Classical Theory of Fields, fourth re-
vised English ed., vol. 2 of Course of Theoretical Physics, Pergamon Press, Ltd.,
Oxford . . . , 1975, ISBN 0-08-025072-6.
[8] D. E. S OPER, Classical Field Theory, John Wiley & Sons, Inc., New York, Lon-
don, Sydney and Toronto, 1976, ISBN 0-471-81368-0.
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
6
Electromagnetic
Fields and
Matter
The microscopic Maxwell equations (1.45) derived in Chapter 1 are valid on all
scales where a classical description is good. However, when macroscopic mat-
ter is present, it is sometimes convenient to use the corresponding macroscopic
Maxwell equations (in a statistical sense) in which auxiliary, derived fields are
introduced in order to incorporate effects of macroscopic matter when this is
immersed fully or partially in an electromagnetic field.
87
i i
i
where the ρ is the charge density introduced in Equation (1.7) on page 5, the
electric dipole moment vector
p(x0 ) = d3x (x − x0 ) ρ(x ) (6.2)
V
1 x − x 1 1
φp (x) = d3x P(x ) ·
=− 3
d x P(x ) · ∇
4πε0 V |x − x |3 4πε0 V |x − x |
1 1
= d3x P(x ) · ∇
4πε0 V |x − x |
(6.5)
Using the expression Equation (M.87) on page 187 and applying the diver-
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
gence theorem, we can rewrite this expression for the potential as follows:
1 3 P(x ) 3 ∇ · P(x )
φp (x) = dx ∇· − dx
4πε0 V |x − x | V |x − x |
(6.6)
1 2 P(x ) · n̂ 3 ∇ · P(x )
= dx − dx
4πε0 S |x − x | V |x − x |
where the first term, which describes the effects of the induced, non-cancelling
dipole moment on the surface of the volume, can be neglected, unless there is
a discontinuity in P · n̂ at the surface. Doing so, we find that the contribution
from the electric dipole moments to the potential is given by
1 −∇ · P(x )
φp = d3x (6.7)
4πε0 V |x − x |
Comparing this expression with expression Equation (3.3) on page 37 for the
electrostatic potential from a static charge distribution ρ, we see that −∇ · P(x)
has the characteristics of a charge density and that, to the lowest order, the
effective charge density becomes ρ(x) − ∇ · P(x), in which the second term is a
polarisation term.
The version of Equation (1.7) on page 5 where free, ‘true’ charges and
bound, polarisation charges are separated thus becomes
D = ε0 E + P (6.9)
we obtain
where ρtrue is the ‘true’ charge density in the medium. This is one of Max-
well’s equations and is valid also for time varying fields. By introducing the
notation ρpol = −∇ · P for the ‘polarised’ charge density in the medium, and
ρtotal = ρtrue + ρpol for the ‘total’ charge density, we can write down the follow-
ing alternative version of Maxwell’s equation (6.23a) on page 92
ρtotal (x)
∇·E = (6.11)
ε0
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
Often, for low enough field strengths |E|, the linear and isotropic relation-
ship between P and E
P = ε0 χE (6.12)
D = εE (6.13)
where, approximately,
ε = ε0 (1 + χ) (6.14)
1. In analogy with ‘true’ charges for the electric case, we may have ‘true’
currents jtrue , i.e., a physical transport of true charges.
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
jM = ∇ × M (6.16)
∂P
jtotal = jtrue + +∇×M (6.17)
∂t
We might then, erroneously, be led to think that
true ∂P
∇ × B = µ0 j + +∇×M (6.18)
∂t
Moving the term ∇ × M to the left hand side and introducing the magnetising
field (magnetic field intensity, Ampère-turn density) as
B
H= −M (6.19)
µ0
and using the definition for D, Equation (6.9) on page 89, we can write this
incorrect equation in the following form
∂P ∂D ∂E
∇ × H = jtrue + = jtrue + − ε0 (6.20)
∂t ∂t ∂t
As we see, in this simplistic view, we would pick up a term which makes
the equation inconsistent; the divergence of the left hand side vanishes while
the divergence of the right hand side does not. Maxwell realised this and to
overcome this inconsistency he was forced to add his famous displacement
current term which precisely compensates for the last term in the right hand
side. In Chapter 1, we discussed an alternative way, based on the postulate of
conservation of electric charge, to introduce the displacement current.
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
µ = µ0 (1 + χm ) (6.22)
∇ · D = ρ(t, x) (6.23a)
∇·B = 0 (6.23b)
∂B
∇×E = − (6.23c)
∂t
∂
∇ × H = j(t, x) + D (6.23d)
∂t
and are called Maxwell’s macroscopic equations. These equations are conveni-
ent to use in certain simple cases. Together with the boundary conditions and
the constitutive relations, they describe uniquely (but only approximately!) the
properties of the electric and magnetic fields in matter.
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
3 j2 3
d x j·E = dx − d3x j · EEMF (6.27)
V V σ V
3 j2 ∂ 3 1
d x j·E EMF
= dx + d3x (E · D + H · B) + d2x (E × H) · n̂
σ ∂t V 2
)V *+ , ) V *+ , ) *+ , )S *+ ,
Applied electric power Joule heat Field energy Radiated power
(6.28)
1
Ue = d3x E · D (6.29)
2 V
1
Um = d3x H · B (6.30)
2 V
S = E×H (6.31)
where U e is the electric field energy, U m is the magnetic field energy, both
measured in J, and S is the Poynting vector (power flux), measured in W/m 2 .
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
One verifies easily that the ith vector components of the two terms in
square brackets in the right hand member of (6.32) can be expressed as
1 ∂D ∂E ∂ 1
[E(∇ · D) − D × (∇ × E)]i = E· −D· + Ei D j − E · D δi j
2 ∂xi ∂xi ∂x j 2
(6.33)
and
1 ∂B ∂H ∂ 1
[H(∇ · B) − B × (∇× H)]i = H· −B· + Hi B j − B · H δi j
2 ∂xi ∂xi ∂x j 2
(6.34)
respectively.
Using these two expressions in the ith component of Equation (6.32) above
and re-shuffling terms, we get
1 ∂D ∂E ∂B ∂H ∂
(ρE + j × B)i − E· −D· + H· −B· + (D × B)i
2 ∂xi ∂xi ∂xi ∂xi ∂t
∂ 1 1
= Ei D j − E · D δi j + Hi B j − H · B δi j
∂x j 2 2
(6.35)
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
D = κε0 E = εE (6.39)
B = κm µ0 H = µH (6.40)
which expresses the balance between the force on the matter, the rate of change
of the electromagnetic field momentum and the Maxwell stress. This equation
is called the momentum theorem in Maxwell’s theory.
In vacuum (6.42) becomes
3 1 d 3
d x ρ(E + v × B) + 2 d x S= d2x T n̂ (6.43)
V c dt V S
or
d mech d field
p + p = d2x T n̂ (6.44)
dt dt S
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
Bibliography
[1] E. H ALLÉN, Electromagnetic Theory, Chapman & Hall, Ltd., London, 1962.
[2] J. D. JACKSON, Classical Electrodynamics, third ed., John Wiley & Sons, Inc.,
New York, NY . . . , 1999, ISBN 0-471-30932-X.
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
7
Electromagnetic
Fields from
Arbitrary Source
Distributions
While, in principle, the electric and magnetic fields can be calculated from the
Maxwell equations in Chapter 1, or even from the wave equations in Chapter 2,
it is often physically more lucid to calculate them from the electromagnetic
potentials derived in Chapter 3. In this chapter we will derive the electric and
magnetic fields from the potentials.
We recall that in order to find the solution (3.31) for the generic inhomo-
geneous wave equation (3.15) on page 41 we presupposed the existence of a
Fourier transform pair (3.16a) on page 41 for the generic source term
∞
Ψ(t, x) = dω Ψω (x) e−iωt (7.1a)
−∞
1 ∞
Ψω (x) = dt Ψ(t, x) eiωt (7.1b)
2π −∞
That such transform pairs exist is true for most physical variables which are
neither strictly monotonically increasing nor strictly monotonically decreasing
with time. For charge and current densities varying in time we can therefore,
without loss of generality, work with individual Fourier components ρ ω (x) and
jω (x), respectively. Strictly speaking, the existence of a single Fourier compon-
ent assumes a monochromatic source (i.e., a source containing only one single
frequency component), which in turn requires that the electric and magnetic
fields exist for infinitely long times. However, by taking the proper limits, we
may still use this approach even for sources and fields of finite duration.
This is the method we shall utilise in this chapter in order to derive the
electric and magnetic fields in vacuum from arbitrary given charge densities
ρ(t, x) and current densities j(t, x), defined by the temporal Fourier transform
97
i i
i
pairs
∞
ρ(t, x) = dω ρω (x) e−iωt (7.2a)
−∞
1 ∞
ρω (x) = dt ρ(t, x) eiωt (7.2b)
2π −∞
and
∞
j(t, x) = dω jω (x) e−iωt (7.3a)
−∞
1 ∞
jω (x) = dt j(t, x) eiωt (7.3b)
2π −∞
under the assumption that only retarded potentials produce physically accept-
able solutions.
The temporal Fourier transform pair for the retarded scalar potential can
then be written
∞
φ(t, x) = dω φω (x) e−iωt (7.4a)
−∞
1 ∞ 1 eik|x−x |
φω (x) = dt φ(t, x) eiωt = d3x ρω (x ) (7.4b)
2π −∞ 4πε0 V |x − x |
where in the last step, we made use of the explicit expression for the temporal
Fourier transform of the generic potential component Ψ ω (x), Equation (3.28)
on page 43. Similarly, the following Fourier transform pair for the vector po-
tential must exist:
∞
A(t, x) = dω Aω (x) e−iωt (7.5a)
−∞
1 ∞ µ0 eik|x−x |
Aω (x) = dt A(t, x) eiωt = d3x jω (x ) (7.5b)
2π −∞ 4π V |x − x |
Clearly, we must require that
Aω = A∗−ω , φω = φ∗−ω (7.6)
in order that all physical quantities be real. Similar transform pairs and re-
quirements of real-valuedness exist for the fields themselves.
In the limit that the sources can be considered monochromatic containing
only one single frequency ω0 , we have the much simpler expressions
ρ(t, x) = ρ0 (x)e−iω0 t (7.7a)
j(t, x) = j0 (x)e−iω0 t (7.7b)
−iω0 t
φ(t, x) = φ0 (x)e (7.7c)
A(t, x) = A0 (x)e−iω0 t (7.7d)
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
The calculations are much simplified if we work in ω space and, at the fi-
nal stage, Fourier transform back to ordinary t space. We are working in the
Lorentz gauge and note that in ω space the Lorentz condition, Equation (3.13)
on page 40, takes the form
k
∇ · Aω − i φω = 0 (7.9)
c
which provides a relation between (the Fourier transforms of) the vector and
scalar potentials.
Using the Fourier transformed version of Equation (7.8) above and Equa-
tion (7.5b) on the facing page, we obtain
µ0 3 eik|x−x |
Bω (x) = ∇ × Aω (x) = ∇ × d x jω (x ) (7.10)
4π V |x − x |
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
where
def ∂j
j̇(tret , x ) ≡ (7.13)
∂t
t=tret
The first term, the induction field, dominates near the current source but falls
off rapidly with distance from it, is the electrodynamic version of the Biot-
Savart law in electrostatics, Formula (1.15) on page 8. The second term, the
radiation field or the far field, dominates at large distances and represents en-
ergy that is transported out to infinity. Note how the spatial derivatives (∇)
gave rise to a time derivative (˙)!
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
i
ρω (x ) = − ∇ · jω (x ) (7.16)
ω
Doing so in the last term of Equation (7.14) above, and also using the fact that
k = ω/c, we can rewrite this Equation as
1 ρω (x )eik|x−x | (x − x )
Eω (x) = d3x
4πε0 V |x − x |3
ik|x−x |
1 3 [∇ · jω (x )](x − x ) e (7.17)
− dx
− ikjω (x )
c V |x − x | |x − x |
) *+ ,
Iω
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
But, since
∂ xl − xl ik|x−x | ∂ jωmxl − xl ik|x−x |
jωm e =
e
∂xm |x − x |2 ∂xm |x − x |2
(7.19)
∂ xl − xl ik|x−x |
+ jωm e
∂xm |x − x |2
we can rewrite Iω as
∂ 3 xl − xl ik|x−x | eik|x−x |
Iω = − d x jωm x̂l e + ikjω
V ∂xm |x − x |2 |x − x |
(7.20)
∂ xl − xl
+ d3x jωm x̂l eik|x−x |
V ∂xm
|x − x | 2
Using the triple product ‘bac-cab’ Formula (F.51) on page 168 backwards, and
inserting the resulting expression for I ω into Equation (7.17) on the previous
page, we arrive at the following final expression for the Fourier transform of
the total E field:
1 eik|x−x | iµ0 ω eik|x−x |
Eω (x) = − ∇ d3x ρω (x ) + d3x jω (x )
4πε0 V |x − x | 4π V |x − x |
|
1 ρω (x )e ik|x−x (x − x )
= d3x
4πε0 V |x − x | 3
1 [jω (x )eik|x−x | · (x − x )](x − x )
3
+ dx (7.22)
c V |x − x |4
ik|x−x | × (x − x )] × (x − x )
1 3 [jω (x )e
+ dx
c V |x − x |4
ik [jω (x )eik|x−x | × (x − x )] × (x − x )
− d3x
c V |x − x |3
Taking the inverse Fourier transform of Equation (7.22), once again using
the vacuum relation ω = kc, we find, at last, the expression in time domain for
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
Here, the first term represents the retarded Coulomb field and the last term
represents the radiation field which carries energy over very large distances.
The other two terms represent an intermediate field which contributes only in
the near zone and must be taken into account there.
With this we have achieved our goal of finding closed-form analytic ex-
pressions for the electric and magnetic fields when the sources of the fields
are completely arbitrary, prescribed distributions of charges and currents. The
only assumption made is that the advanced potentials have been discarded;
recall the discussion following Equation (3.31) on page 43 in Chapter 3.
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
page, which give the total electric and magnetic fields, we obtain
∞ , x ) × (x − x )
µ0 j̇(tret
Brad (t, x) = dω Brad ω (x) e−iωt = d3x
−∞ 4πc V |x − x |2
(7.24a)
∞
Erad (t, x) = dω Erad ω (x) e−iωt
−∞
, x ) × (x − x )] × (x − x ) (7.24b)
1 3 [j̇(tret
= dx
4πε0 c2 V |x − x |3
where
def ∂j
j̇(tret , x ) ≡ (7.25)
∂t
t=tret
Instead of studying the fields in the time domain, we can often make a
spectrum analysis into the frequency domain and study each Fourier compon-
ent separately. A superposition of all these components and a transformation
back to the time domain will then yield the complete solution.
The Fourier representation of the radiation fields Equation (7.24a) and
Equation (7.24b) above were included in Equation (7.11) on page 100 and
Equation (7.22) on page 102, respectively and are explicitly given by
1 ∞
ω (x) =
Brad dt Brad (t, x) eiωt
2π −∞
kµ0 jω (x ) × (x − x ) ik|x−x |
= −i d3x e (7.26a)
4π V |x − x |2
µ0 jω (x ) × k ik|x−x |
= −i d3x e
4π V |x − x |
1 ∞
Erad
ω (x) = dt Erad (t, x) eiωt
2π −∞
k [jω (x ) × (x − x )] × (x − x ) ik|x−x |
= −i d3x e (7.26b)
4πε0 c V |x − x |3
1 [jω (x ) × k] × (x − x ) ik|x−x |
= −i d3x e
4πε0 c V |x − x |2
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
dS = n̂d2x
x − x
k̂ x − x0 x x − x0
x x0
V
F IGURE 7.1: Relation between the surface normal and the k vector for
radiation generated at source points x near the point x0 in the source
volume V. At distances much larger than the extent of V, the unit vector
n̂, normal to the surface S which has its centre at x0 , and the unit vector
k̂ of the radiation k vector from x are nearly coincident.
on x0 , has a large enough radius |x − x0 | max |x − x0 |, we see from Figure 7.1
that we can approximate
k x − x ≡ k · (x − x ) ≡ k · (x − x0 ) − k · (x − x0 )
(7.27)
≈ k |x − x0 | − k · (x − x0 )
Recalling from Formula (F.45) and Formula (F.46) on page 168 that
dS = |x − x0 |2 dΩ = |x − x0 |2 sin θ dθ dϕ
and noting from Figure 7.1 that k̂ and n̂ are nearly parallel, we see that we can
approximate.
k̂ · dS k̂ · n̂
= dS ≈ dΩ (7.28)
|x − x0 | 2
|x − x0 |2
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
1 1
S
= E × H
= Re E × B∗ = Re Eω e−iωt × (Bω e−iωt )∗
2µ0 2µ0
(7.30)
1 1
= Re Eω × B∗ω e−iωt eiωt = Re Eω × B∗ω
2µ0 2µ0
Using the far-field approximations (7.29a) and (7.29b) and the fact that 1/c =
√ √
ε0 µ0 and R0 = µ0 /ε0 according to the definition (2.18) on page 28, we ob-
tain
2
1 1
−ik·(x −x0 ) x − x0
S
= R d 3
x (j × k)e (7.31)
32π2 |x − x0 |2 V |x − x0 |
0 ω
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
If we carry out the temporal integration first and use the fact that
∞
dt e−i(ω+ω )t = 2πδ(ω + ω ) (7.34)
−∞
where the last step follows from the real-valuedness of E ω and Bω . We in-
sert the Fourier transforms of the field components which dominate at large
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
distances, i.e., the radiation fields (7.26a) and (7.26b). The result, after integ-
ration over the area S of a large sphere which encloses the source, is
∞ 2
1 µ0 j ω × k |
U= 2
d x n̂· dω d x 3
e ik|x−x k̂
(7.36)
4π ε0 S 0 V |x − x |
Inserting the approximations (7.27) and (7.28) into Equation (7.36) above
and also introducing
∞
U= Uω dω (7.37)
0
and recalling the definition (2.18) on page 28 for the vacuum resistance R 0 we
obtain
2
dUω 1
3
dω ≈ R0 d x (jω × k)e −ik·(x −x0 )
dΩ 4π V
dω (7.38)
Bibliography
[1] F. H OYLE , S IR AND J. V. NARLIKAR, Lectures on Cosmology and Action at a
Distance Electrodynamics, World Scientific Publishing Co. Pte. Ltd, Singapore,
New Jersey, London and Hong Kong, 1996, ISBN 9810-02-2573-3(pbk).
[2] J. D. JACKSON, Classical Electrodynamics, third ed., John Wiley & Sons, Inc.,
New York, NY . . . , 1999, ISBN 0-471-30932-X.
[3] L. D. L ANDAU AND E. M. L IFSHITZ, The Classical Theory of Fields, fourth re-
vised English ed., vol. 2 of Course of Theoretical Physics, Pergamon Press, Ltd.,
Oxford . . . , 1975, ISBN 0-08-025072-6.
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
8
Electromagnetic
Radiation and
Radiating
Systems
In Chapter 3 we were able to derive general expressions for the scalar and vec-
tor potentials from which we then, in Chapter 7, calculated the total electric
and magnetic fields from arbitrary distributions of charge and current sources.
The only limitation in the calculation of the fields was that the advanced po-
tentials were discarded.
Thus, one can, at least in principle, calculate the radiated fields, Poynting
flux and energy for an arbitrary current density Fourier component and then
add these Fourier components together to construct the complete electromag-
netic field at any time at any point in space. However, in practice, it is often
difficult to evaluate the source integrals unless the current has a simple distri-
bution in space. In the general case, one has to resort to approximations. We
shall consider both these situations.
109
i i
i
sin[k(L/2 − x3 )]
− L2 j(t , x ) L
2
F IGURE 8.1: A linear antenna used for transmission. The current in the
feeder and the antenna wire is set up by the EMF of the generator (the
transmitter). At the ends of the wire, the current is reflected back with a
180◦ phase shift to produce a antenna current in the form of a standing
wave.
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
r̂
x3 = z ϕ̂
x
L
2
θ̂
θ
k̂
jω (x )
x2
x1
− L2
In order to evaluate Formula (7.32) on page 107 with the explicit mono-
chromatic current (8.1) inserted, we use a spherical polar coordinate system as
in Figure 8.2 to evaluate the source integral
2
−x )
d x j0 × k e
3 −ik·(x 0
V
2
L/2 sin[k(L/2 − x )]
−ikx3 cos θ ikx0 cos θ
= I0 3
k sin θe e dx3
−L/2 sin(kL/2)
2
2 2
2 k sin θ ikx0 cos θ 2
L/2
= I0 2 e 2 sin[k(L/2 − x3 )] cos(kx3 cos θ) dx3
sin (kL/2) 0
cos[(kL/2) cosθ] − cos(kL/2) 2
= 4I02
sin θ sin(kL/2)
(8.2)
Inserting this expression and dΩ = 2π sin θ dθ into Formula (7.32) on page 107
and integrating over θ, we find that the total radiated power from the antenna
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
is
π
1 cos[(kL/2) cosθ] − cos(kL/2) 2
P(L) = R0 I02 sin θ dθ (8.3)
4π 0 sin θ sin(kL/2)
One can show that
2
π L
lim P(L) = R0 I02 (8.4)
kL→0 12 λ
where λ is the vacuum wavelength.
The quantity
2 2
P(L) P(L) π L L
R (L) = 2 = 1 2 = R0
rad
≈ 197 Ω (8.5)
Ieff 2 I0
6 λ λ
where in the last step the Euler-Mascheroni constant γ = 0.5772 . . . and the
cosine integral Ci(x) were introduced. Inserting this into the expression Equa-
tion (8.6) we obtain the value Rrad (λ/2) ≈ 73 Ω.
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
r̂
x3 = z = z ϕ̂
x
θ̂
θ
k̂
x2
ẑ
jω (x ) ϕ
x
x1 ϕ̂
ϕ
ρ̂
F IGURE 8.3: For the loop antenna the spherical coordinate system
(r, θ, ϕ) describes the field point x (the radiation field) and the cylindrical
coordinate system (ρ , ϕ , z ) describes the source point x (the antenna
current).
In our case the generator produces a single frequency ω and we feed the an-
tenna across a small gap where the loop crosses the positive x 1 axis. The cir-
cumference of the loop is chosen to be exactly one wavelength λ = 2πc/ω. This
means that the antenna current oscillates in the form of a sinusoidal standing
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
Utilising the periodicity of the integrands over the integration interval [0, 2π],
introducing the auxiliary integration variable ϕ = ϕ −ϕ, and utilising standard
trigonometric identities, the first integral in the RHS of (8.15) can be rewritten
2π
e−ika sin θ cos ϕ cos ϕ cos(ϕ + ϕ) dϕ
0
2π
= cos ϕ e−ika sin θ cos ϕ cos2 ϕ dϕ + a vanishing integral
0
2π
−ika sin θ cos ϕ 1 1
= cos ϕ e + cos 2ϕ dϕ (8.16)
0 2 2
2π
1
= cos ϕ e−ika sin θ cos ϕ dϕ
2 0
2π
1
+ cos ϕ e−ika sin θ cos ϕ cos 2ϕ dϕ
2 0
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
and compare with Equation (8.23) on the preceding page, we see that π(t, x)
satisfies this equation of continuity. Furthermore, if we compare with the elec-
tric polarisation [cf. Equation (6.9) on page 89], we see that the quantity π
is related to the “true” charges in the same way as P is related to polarised
charge, namely as a dipole moment density. The quantity π is referred to as
the polarisation vector since, formally, it treats also the “true” (free) charges
as polarisation charges so that
ρtrue + ρpol −∇ · π − ∇ · P
∇·E = = (8.25)
ε0 ε0
∇ · Πe = −φ (8.26a)
1 ∂Πe
=A (8.26b)
c2 ∂t
where φ and A are the electromagnetic scalar and vector potentials, respect-
ively. As we see, Πe acts as a “super-potential” in the sense that it is a poten-
tial from which we can obtain other potentials. It is called the Hertz’ vector
or polarisation potential. Requiring that the scalar and vector potentials φ and
A, respectively, fulfil their inhomogeneous wave equations, one finds, using
(8.24) and (8.26), that Hertz’ vector must satisfy the inhomogeneous wave
equation
1 ∂2 e π
2 Πe = Π − ∇2 Πe = (8.27)
c ∂t
2 2 ε0
This equation is of the same type as Equation (3.15) on page 41, and has
therefore the retarded solution
, x )
1 π(tret
Πe (t, x) = d3x (8.28)
4πε0 V |x − x |
C = ∇ × Πe (8.30)
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
x − x
x − x0 x
Θ
x x0
V
we see that we can calculate the magnetic and electric fields, respectively, as
follows
1 ∂C
B= (8.31a)
c2 ∂t
E = ∇×C (8.31b)
Clearly, the last equation is valid only outside the source volume, where
∇ · E = 0. Since we are mainly interested in the fields in the far zone, a long
distance from the source region, this is no essential limitation.
Assume that the source region is a limited volume around some central
point x0 far away from the field (observation) point x illustrated in Figure 8.4.
Under these assumptions, we can expand the Hertz’ vector, expression (8.28)
on the previous page, due to the presence of non-vanishing π(t ret , x ) in the
eik|x−x | eik|(x−x0 )−(x −x0 )|
≡
|x − x | |(x − x0 ) − (x − x0 )|
∞ (8.32)
= ik ∑ (2n + 1)Pn (cos Θ) jn (k x − x0 )h(1)
n (k |x − x0 |)
n=0
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
where
eik|x−x |
is a Green function
|x − x |
Θ is the angle between x − x0 and x − x0 (see Figure 8.4 on page 118)
Pn (cos Θ) is the Legendre polynomial of order n
jn (k x − x0 ) is the spherical Bessel function of the first kind of order n
n (k |x − x0 |) is the spherical Hankel function of the first kind of order n
h(1)
Inserting Equation (8.32) on the facing page, together with Equation (8.33)
above, into Equation (8.29) on page 117, we can in a formally exact way ex-
pand the Fourier component of the Hertz’ vector as
ik ∞ n
Πeω = ∑ ∑ (2n + 1)(−1)mh(1)
4πε0 n=0 n (k |x − x0 |) Pn (cos θ) e
m imϕ
m=−n (8.35)
× d3x πω (x ) jn (k x − x0 ) P−m −imϕ
n (cos θ ) e
V
eik|x−x0 |
n (k |x − x0 |) ≈ (−i)
h(1) n+1
(8.37)
k |x − x0 |
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
and replace jn with the first term in its power series expansion:
2n n!
n
jn (k x − x0 ) ≈ k x − x0 (8.38)
(2n + 1)!
Inserting these expansions into Equation (8.35) on the previous page, we obtain
the multipole expansion of the Fourier component of the Hertz’ vector
∞
Πeω ≈ ∑ Πeω(n) (8.39a)
n=0
where
1 eik|x−x0 | 2n n!
Πeω(n) = (−i)n d3x πω (x ) (k x − x0 )n Pn (cos Θ) (8.39b)
4πε0 |x − x0 | (2n)! V
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
k̂
x3
x Brad
Erad
θ
p r̂
x2
x1
[cf. Equation (6.2) on page 88 which describes the static dipole moment]. If
a spherical coordinate system is chosen with its polar axis along p ω as in Fig-
ure 8.5, the components of Πeω(0) are
def 1 eik|x−x0 |
Πer ≡ Πeω(0) · r̂ = pω cos θ (8.42a)
4πε0 |x − x0 |
def 1 eik|x−x0 |
Πeθ ≡ Πeω(0) · θ̂ = − pω sin θ (8.42b)
4πε0 |x − x0 |
def
Πeϕ ≡ Πeω(0) · ϕ̂ = 0 (8.42c)
Evaluating Formula (8.30) on page 117 for the help vector C, with the
spherically polar components (8.42) of Π eω(0) inserted, we obtain
ik|x−x0 |
1 1 e
Cω = Cω,ϕ
(0)
ϕ̂ = − ik pω sin θ ϕ̂ (8.43)
4πε0 |x − x0 | |x − x0 |
Applying this to Equation (8.31) on page 118, we obtain directly the Fourier
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
Keeping only those parts of the fields which dominate at large distances
(the radiation fields) and recalling that the wave vector k = k(x − x 0 )/ |x − x0 |
where k = ω/c, we can now write down the Fourier components of the radiation
parts of the magnetic and electric fields from the dipole:
These fields constitute the electric dipole radiation, also known as E1 radi-
ation.
and introducing
ηi = xi − x0,i (8.48a)
ηi = xi − x0,i (8.48b)
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
The final result is that the antisymmetric, magnetic dipole, part of Π eω(1)
can be written
(1) k eik|x−x0 |
Πe,antisym
ω =− (x − x0 ) × mω (8.55)
4πε0 ω |x − x0 |2
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
In analogy with the electric dipole case, we insert this expression into Equa-
tion (8.30) on page 117 to evaluate C, with which Equations (8.31) on page 118
then gives the B and E fields. Discarding, as before, all terms belonging to the
near fields and transition fields and keeping only the terms that dominate at
large distances, we obtain
µ0 eik|x−x0 |
ω (x) = −
Brad (mω × k) × k (8.56a)
4π |x − x0 |
k eik|x−x0 |
ω (x) =
Erad mω × k (8.56b)
4πε0 c |x − x0 |
which are the fields of the magnetic dipole radiation (M1 radiation).
Again we use this expression in Equation (8.30) on page 117 to calculate the
fields via Equations (8.31) on page 118. Tedious, but fairly straightforward
algebra (which we will not present here), yields the resulting fields. The ra-
diation components of the fields in the far field zone (wave zone) are given
by
iµ0 ω eik|x−x0 |
ω (x) =
Brad k · Qω × k (8.58a)
8π |x − x0 |
i eik|x−x0 |
Erad
ω (x) = k · Qω × k × k (8.58b)
8πε0 |x − x0 |
This type of radiation is called electric quadrupole radiation or E2 radiation.
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
studied above. In order to handle this non-stationary situation, we use the re-
tarded potentials (3.32) on page 44 in Chapter 3
1 ρ(t , x )
φ(t, x) = d3x ret (8.59a)
4πε0 V |x − x |
, x )
µ0 j(t
A(t, x) = d3x ret (8.59b)
4π V |x − x |
and consider a source region with such a limited spatial extent that the charges
and currents are well localised. Specifically, we consider a charge q , for in-
stance an electron, which, classically, can be thought of as a localised, unstruc-
tured and rigid “charge distribution” with a small, finite radius. The part of this
“charge distribution” dq which we are considering is located in dV = d3x in
the sphere in Figure 8.6 on the next page. Since we assume that the electron
(or any other other similar electric charge) is moving with a velocity v whose
direction is arbitrary and whose magnitude can be almost comparable to the
speed
of light, we cannot
say that the charge and current to be used in (8.59) is
ρ(t , x ) d3x and 3
V ret V vρ(t ret , x ) d x , respectively, because in the finite time
interval during which the observed signal is generated, part of the charge dis-
tribution will “leak” out of the volume element d 3x .
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
1111111111
0000000000
x(t) dr
0000
1111
v(t )
0000000000
1111111111 dS
0000
1111
x − x
0000000000
1111111111
dV
x (t ) q
F IGURE 8.6: Signals which are observed at the field point x at time t were
generated at source points x (t ) on a sphere, centred on x and expanding,
as time increases, with the velocity c outward from the centre. The source
charge element moves with an arbitrary velocity v and gives rise to a
source “leakage” out of the source volume dV = d3x .
This is the expression to be used in the Formulae (8.59) on the preceding page
for the retarded potentials. The result is (recall that j = ρv)
1 dq
φ(t, x) = (8.64a)
4πε0 |x − x | − (x−xc )·v
µ0 v dq
A(t, x) = (8.64b)
4π |x − x | − (x−xc )·v
For a sufficiently small and well localised charge distribution we can, assuming
that the integrands do not change sign in the integration volume, use the mean
value theorem and the fact that V dq = q to evaluate these expressions to
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
become
q 1 q 1
φ(t, x) = = (8.65a)
4πε0 |x − x | − (x−xc )·v 4πε0 s
q v q v v
A(t, x) = = = φ(t, x) (8.65b)
4πε0 c2 |x − x | − (x−xc )·v 4πε0 c2 s c2
where
(x − x (t )) · v(t )
s = s(t , x) = x − x (t ) − (8.66a)
c
x − x (t ) v(t )
= x − x (t ) 1 −
· (8.66b)
|x − x (t )| c
x − x (t ) v(t )
= (x − x (t )) · − (8.66c)
|x − x (t )| c
is the retarded relative distance. The potentials (8.65) are precisely the Liénard-
Wiechert potentials which we derived in Section 4.3.2 on page 62 by using a
covariant formalism.
It is important to realise that in the complicated derivation presented here,
the observer is in a coordinate system which has an “absolute” meaning and the
velocity v is that of the particle, whereas in the covariant derivation two frames
of equal standing were moving relative to each other with v. Expressed in the
four-potential, Equation (4.48) on page 61, the Liénard-Wiechert potentials
become
µ κ q 1 v φ
A (x ) = , = ,A (8.67)
4πε0 cs c2 s c
The Liénard-Wiechert potentials are applicable to all problems where a
spatially localised charge emits electromagnetic radiation, and we shall now
study such emission problems. The electric and magnetic fields are calculated
from the potentials in the usual way:
B(t, x) = ∇ × A(t, x) (8.68a)
∂A(t, x)
E(t, x) = −∇φ(t, x) − (8.68b)
∂t
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
|x − x |
q v x0 (t) v(t )
c
x (t ) θ θ0
x − x0
x − x
x(t)
F IGURE 8.7: Signals which are observed at the field point x at time t
were generated at the source point x (t ). After time t the particle, which
moves with nonuniform velocity, has followed a yet unknown trajectory.
Extrapolating tangentially the trajectory from x (t ), based on the velo-
city v(t ), defines the virtual simultaneous coordinate x0 (t).
(in the interest of simplifying our notation, we drop the subscript “ret” on t
from now on). This means that we know the trajectory of the charge q , i.e., x ,
for all times up to the time t at which a signal was emitted in order to precisely
arrive at the field point x at time t. Because of the finite speed of propagation
of the fields, the trajectory at times later than t is not (yet) known.
The retarded velocity and acceleration at time t are given by
dx
v(t ) = (8.70a)
dt
dv d2 x
a(t ) = v̇(t ) = = (8.70b)
dt dt 2
As for the charge coordinate x itself, we have in general no knowledge of the
velocity and acceleration at times later than t , in particular not at the time of
observation t. If we choose the field point x as fixed, application of (8.70) to
the relative vector x − x yields
d
(x − x (t )) = −v(t ) (8.71a)
dt
d2
(x − x (t )) = −v̇(t ) (8.71b)
dt 2
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
The retarded time t can, at least in principle, be calculated from the implicit
relation
|x − x (t )|
t = t (t, x) = t − (8.72)
c
and we shall see later how this relation can be taken into account in the calcu-
lations.
According to Formulae (8.68) on page 127 the electric and magnetic fields
are determined via differentiation of the retarded potentials at the observation
time t and at the observation point x. In these formulae the unprimed ∇, i.e.,
the spatial derivative differentiation operator ∇ = x̂ i ∂/∂xi means that we differ-
entiate with respect to the coordinates x = (x 1 , x2 , x3 ) while keeping t fixed, and
the unprimed time derivative operator ∂/∂t means that we differentiate with re-
spect to t while keeping x fixed. But the Liénard-Wiechert potentials φ and A,
Equations (8.65) on page 127, are expressed in the charge velocity v(t ) given
by Equation (8.70a) on the preceding page and the retarded relative distance
s(t , x) given by Equation (8.66) on page 127. This means that the expressions
for the potentials φ and A contain terms which are expressed explicitly in t ,
which in turn is expressed implicitly in t via Equation (8.72) above. Despite
this complication it is possible, as we shall see below, to determine the electric
and magnetic fields and associated quantities at the time of observation t. To
this end, we need to investigate carefully the action of differentiation on the
potentials.
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
x − x
(∇)t t = − (8.78)
cs
which gives the following operator relation when (∇) t is acting on an arbitrary
function of t and x:
∂ x − x ∂
(∇)t = (∇)t t + (∇)t = −
+ (∇)t (8.79)
∂t x cs ∂t x
With the help of the rules (8.79) and (8.76) we are now able to replace t by t
in the operations which we need to perform. We find, for instance, that
1 q
∇φ ≡ (∇φ)t = ∇
4πε0 s
(8.80a)
q x − x v(t ) x − x ∂s
=− − −
4πε0 s2 |x − x | c cs ∂t x
∂A ∂A ∂ µ0 q v(t )
≡ =
∂t ∂t x ∂t 4π s
x
(8.80b)
q ∂s
= x − x sv̇(t ) − x − x v(t )
4πε0 c2 s3 ∂t x
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
Utilising these relations in the calculation of the E field from the Liénard-
Wiechert potentials, Equations (8.65) on page 127, we obtain
∂
E(t, x) = −∇φ(t, x) − A(t, x)
∂t
q (x − x (t )) − |x − x (t )| v(t )/c
=
4πε0 s2 (t , x) |x − x (t )|
(x − x (t )) − |x − x (t )| v(t )/c ∂s(t , x) |x − x (t )| v̇(t )
− −
cs(t , x) ∂t x c2
(8.81)
Starting from expression (8.66a) on page 127 for the retarded relative distance
s(t , x), we see that we can evaluate (∂s/∂t )x in the following way
∂s ∂
(x − x ) · v(t )
= x−x −
∂t ∂t c
x x
1
2
∂
1 ∂ x − x (t ) ∂v(t )
= x − x (t ) − · v(t ) + (x − x (t )) ·
∂t c ∂t ∂t
(x − x ) · v(t ) v2 (t ) (x − x ) · v̇(t )
=− + −
|x − x | c c
(8.82)
where Equation (8.73) on page 129 and Equations (8.70) on page 128, respect-
ively, were used. Hence, the electric field generated by an arbitrarily moving
charged particle at x (t ) is given by the expression
q |x − x (t )| v(t ) v2 (t )
E(t, x) = (x − x (t )) − 1 −
4πε0 s3 (t , x) c c2
) *+ ,
Coulomb field when v → 0
q x − x (t ) |x − x (t )| v(t )
+ × (x − x (t )) − × v̇(t )
4πε0 s3 (t , x) c2 c
) *+ ,
Radiation (acceleration) field
(8.83)
The first part of the field, the velocity field, tends to the ordinary Coulomb field
when v → 0 and does not contribute to the radiation. The second part of the
field, the acceleration field, is radiated into the far zone and is therefore also
called the radiation field.
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
From Figure 8.7 on page 128 we see that the position the charged particle
would have had if at t all external forces would have been switched off so that
the trajectory from then on would have been a straight line in the direction of
the tangent at x (t ) is x0 (t), the virtual simultaneous coordinate. During the
arbitrary motion, we interpret x − x 0 as the coordinate of the field point x relat-
ive to the virtual simultaneous coordinate x 0 (t). Since the time it takes from a
signal to propagate (in the assumed vacuum) from x (t ) to x is |x − x | /c, this
relative vector is given by
|x − x (t )| v(t )
x − x0 (t) = x − x (t ) − (8.84)
c
This allows us to rewrite Equation (8.83) on the previous page in the following
way
q v2 (x − x0 ) × v̇
E(t, x) = (x − x0 ) 1 − 2 + (x − x ) × (8.85)
4πε0 s3 c c2
In a similar manner we can compute the magnetic field:
x − x ∂
B(t, x) = ∇ × A(t, x) ≡ (∇)t × A = (∇)t × A − × A
cs ∂t x
(8.86)
q x − x x − x ∂A
=−
×v− ×
4πε0 c s |x − x |
2 2 c |x − x | ∂t x
where we made use of Equation (8.65) on page 127 and Formula (8.76) on
page 130. But, according to (8.80a),
x − x q x − x
× (∇) t φ = ×v (8.87)
c |x − x | 4πε0 c2 s2 |x − x |
so that
x − x ∂A
B(t, x) = × −(∇φ) t −
c |x − x | ∂t x
(8.88)
x−x
= × E(t, x)
c |x − x |
The radiation part of the electric field is obtained from the acceleration
field in Formula (8.83) on the preceding page as
Erad (t, x) = lim E(t, x)
|x−x |→∞
q |x − x | v
= (x − x ) × (x − x ) − × v̇ (8.89)
4πε0 c2 s3 c
q
= (x − x ) × [(x − x0 ) × v̇]
4πε0 c2 s3
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
where in the last step we again used Formula (8.84) on the preceding page.
Using this formula and Formula (8.88) on the facing page, the radiation part of
the magnetic field can be written
x − x
Brad (t, x) = × Erad (t, x) (8.90)
c |x − x |
Furthermore, from Equation (8.84) on the facing page, we obtain the following
identity:
which, when inserted into Equation (8.93) above, yields the relation
2
(x − x ) · v |x − x |2 v2 (x − x0 ) × v 2
= − (8.95)
c c2 c
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
Inserting the above into expression (8.91) on the preceding page for s 2 , this
expression becomes
2
2 (x − x ) · v |x − x |2 v2 (x − x0 ) × v
s2 = x − x − 2 x − x + −
c c2 c
|x − x | v
2 2
(x − x0 ) × v
= (x − x ) − −
c c
2 (8.96)
(x − x 0 ) × v
= (x − x0 )2 −
c
(x − x0 (t)) × v(t ) 2
≡ |x − x0 (t)| −
2
c
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
obtain
& 2 '
v2 (x − x0 ) · v
∇s = ∇ |x − x0 | 1 − 2 +
2 2
c c
v2 vv (8.98)
= 2 (x − x0 ) 1 − 2 + 2 · (x − x0 )
c c
v v
= 2 (x − x0 ) + × × (x − x0 )
c c
which we shall use in the following example of a uniformly, unaccelerated
motion of the charge.
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
q 1 q
∇φ = ∇ =− ∇s2
4πε0 s 8πε0 s3
(8.101)
q v v
=− (x − x0 ) + × × (x − x0 )
4πε0 s3 c c
When this expression for ∇φ is inserted into Equation (8.100a) on the preceding page,
the following result
vv q vv
E(t, x) = 2 − 1 · ∇φ = − − 1 · ∇s2
c 8πε0 s3 c2
q v v
= (x − x 0 ) + × × (x − x 0 )
4πε0 s3 c c
v v vv v v
− · (x − x0 ) − 2 · × × (x − x0 )
c c c c c
(8.102)
q v v v2
= (x − x0 ) + · (x − x0 ) − (x − x0 ) 2
4πε0 s3 c c c
v v
− · (x − x0 )
c c
q v2
= (x − x0 ) 1 −
4πε0 s3 c2
follows. Of course, the same result also follows from Equation (8.85) on page 132
with v̇ ≡ 0 inserted.
From Equation (8.102) we conclude that E is directed along the vector from the sim-
ultaneous coordinate x0 (t) to the field (observation) coordinate x(t). In a similar way,
the magnetic field can be calculated and one finds that
µ0 q v2 1
B(t, x) = 3
1 − 2 v × (x − x0 ) = 2 v × E (8.103)
4πs c c
From these explicit formulae for the E and B fields we can discern the following cases:
3. v → c ⇒ E becomes dependent on θ0
E ND OF EXAMPLE 8.1
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
For a localised charge where ρ d3x = q , these expressions reduce to
q
φ(t, x) = (8.112a)
4πε0 s
q v
A(t, x) = (8.112b)
4πε0 c2 s
which we recognise as the Liénard-Wiechert potentials; cf. Equations (8.65) on
page 127. We notice, however, that the derivation here, based on a mathematical
technique which in fact is a Lorentz transformation, is of more general validity than
the one leading to Equations (8.65) on page 127.
Let us now consider the action of the fields produced from a moving, rigid charge
distribution represented by q moving with velocity v, on a charged particle q, also
moving with velocity v. This force is given by the Lorentz force
F = q(E + v × B) (8.113)
With the help of Equation (8.103) on page 136 and Equations (8.111) on the previous
page, and the fact that ∂t = −v · ∇ [cf.. Formula (8.99) on page 135], we can rewrite
expression (8.113) above as
v v v v v
F = q E+v× 2 ×E = q · ∇φ − ∇φ − × × ∇φ (8.114)
c c c c c
Applying the “bac-cab” rule, Formula (F.51) on page 168, on the last term yields
v v v v v2
× × ∇φ = · ∇φ − ∇φ (8.115)
c c c c c2
which means that we can write
F = −q∇ψ (8.116)
where
v2
ψ = 1− 2 φ (8.117)
c
The scalar function ψ is called the convection potential or the Heaviside potential.
When the rigid charge distribution is well localised so that we can use the potentials
(8.112) the convection potential becomes
v2 q
ψ = 1− 2 (8.118)
c 4πε0 s
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
q v̇ = p̈ → −ω2 pω (8.125a)
x − x = x − x0 (8.125b)
The power flux in the far zone is described by the Poynting vector as a
function of Erad and Brad . We use the close correspondence with the dipole
case to find that it becomes
µ0 q 2 (v̇)2 2 x−x
S= sin θ (8.126)
16π2 c |x − x |2 |x − x |
where θ is the angle between v̇ and x − x0 . The total radiated power (integrated
over a closed spherical surface) becomes
µ0 q 2 (v̇)2 q 2 v̇2
P= = (8.127)
6πc 6πε0 c3
which is the Larmor formula for radiated power from an accelerated charge.
Note that here we are treating a charge with v c but otherwise totally un-
specified motion while we compare with formulae derived for a stationary os-
cillating dipole. The electric and magnetic fields, Equation (8.122) on the
previous page and Equation (8.123) on the preceding page, respectively, and
the expressions for the Poynting flux and power derived from them, are here
instantaneous values, dependent on the instantaneous position of the charge
at x (t ). The angular distribution is that which is “frozen” to the point from
which the energy is radiated.
8.3.3 Bremsstrahlung
An important special case of radiation is when the velocity v and the acceler-
ation v̇ are collinear (parallel or anti-parallel) so that v × v̇ = 0. This condition
(for an arbitrary magnitude of v) inserted into expression (8.89) on page 132
for the radiation field, yields
q
Erad (t, x) = (x − x ) × [(x − x ) × v̇], v v̇ (8.128)
4πε0 c2 s3
from which we obtain, with the use of Formula (8.88) on page 132, the mag-
netic field
q |x − x |
Brad (t, x) = [v̇ × (x − x )], v v̇ (8.129)
4πε0 c3 s3
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
v = 0.5c
v = 0.25c
v=0
v
F IGURE 8.8: Polar diagram of the energy loss angular distribution factor
sin2 θ/(1 − v cos θ/c)5 during bremsstrahlung for particle speeds v = 0, v =
0.25c, and v = 0.5c.
The difference between this case and the previous case of v c is that the
approximate expression (8.120) on page 139 for s is no longer valid; we must
instead use the correct expression (8.66) on page 127. The angular distribution
of the power flux (Poynting vector) therefore becomes
µ0 q 2 v̇2 sin2 θ x − x
S=
(8.130)
16π2 c |x − x |2 1 − v cos θ 6 |x − x |
c
It is interesting to note that the magnitudes of the electric and magnetic fields
are the same whether v and v̇ are parallel or anti-parallel.
We must be careful when we compute the energy (S integrated over time).
The Poynting vector is related to the time t when it is measured and to a fixed
surface in space. The radiated power into a solid angle element dΩ, measured
relative to the particle’s retarded position, is given by the formula
On the other hand, the radiation loss due to radiation from the charge at re-
tarded time t :
dU rad dU rad ∂t
dΩ = dΩ (8.132)
dt dt ∂t x
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
dS
dr
x dΩ
x − x + c dt
θ q 2
x2 vdt x1
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
From Figure 8.9 we see that the volume element subtending the solid angle
element
dS
dΩ = (8.135)
x − x 2
2
is
2
d3x = dS dr = x − x2 dΩ dr (8.136)
Here, dr denotes the differential distance between the two spheres and can be
evaluated in the following way
x − x2
dr = x − x2 + c dt − x − x2 − · v dt
x − x2
) *+ ,
v cos θ (8.137)
1 2
x − x2
cs
= c − · v dt = dt
x − x2 x − x2
where Formula (8.66) on page 127 was used in the last step. Hence, the volume
element under consideration is
s
d3x = dS dr = dS cdt (8.138)
x − x 2
We see that the energy which is radiated per unit solid angle during the time
interval (t , t + dt ) is located in a volume element whose size is θ depend-
ent. This explains the difference between expression (8.131) on page 141 and
expression (8.134) on the facing page.
Let the radiated energy, integrated over Ω, be denoted Ũ rad . After tedious,
but relatively straightforward integration of Formula (8.134) on the preceding
page, one obtains
−3
dŨ rad µ0 q 2 v̇2 1 2 q 2 v̇2 v2
= 3 = 1 − (8.139)
dt 6πc v2 3 4πε0 c3 c2
1− c2
If we know v(t ), we can integrate this expression over t and obtain the total
energy radiated during the acceleration or deceleration of the particle. This
way we obtain a classical picture of bremsstrahlung (braking radiation). Of-
ten, an atomistic treatment is required for an acceptable result.
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
E XAMPLE 8.3 B REMSSTRAHLUNG FOR LOW SPEEDS AND SHORT ACCELERATION TIMES
Calculate the bremsstrahlung when a charged particle, moving at a non-relativistic
speed, is accelerated or decelerated during an infinitely short time interval.
We approximate the velocity change at time t = t0 by a delta function:
v̇(t ) = ∆v δ(t − t0 ) (8.140)
which means that
∞
∆v(t0 ) = v̇ dt (8.141)
−∞
Also, we assume v/c 1 so that, according to Formula (8.66) on page 127,
s ≈ x − x (8.142)
and, according to Formula (8.84) on page 132,
x − x0 ≈ x − x (8.143)
According to Parseval’s identity [cf. Equation (7.35) on page 107] the following equal-
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
ity holds:
∞ ∞
E 2 dt = 4π |Eω |2 dω (8.148)
−∞ 0
which means that the radiated energy in the frequency interval (ω, ω + dω) is
Ũω dω = 4πε0 c
rad
|Eω | d x dω
2 2
(8.149)
S
For our infinite spectrum, Equation (8.146) on the facing page, we obtain
q 2 (∆v)2 sin2 θ 2
Ũωrad dω = d x dω
16π3 ε0 c3 S |x − x |2
2π π
q 2 (∆v)2
= dϕ sin2 θ sin θ dθ dω (8.150)
16π3 ε0 c3 0 0
2 2
q ∆v dω
=
3πε0 c c 2π
We see that the energy spectrum Ũωrad is independent of frequency ω. This means that
if we would integrate it over all frequencies ω ∈ [0, ∞], a divergent integral would
result.
In reality, all spectra have finite widths, with an upper cutoff limit set by the quantum
condition
1
h̄ωmax = m(∆v)2 (8.151)
2
which expresses that the highest possible frequency ωmax in the spectrum is that for
which all kinetic energy difference has gone into one single field quantum (photon)
with energy h̄ωmax . If we adopt the picture that the total energy is quantised in terms
of Nω photons radiated during the process, we find that
Ũωrad dω
= dNω (8.152)
h̄ω
or, for an electron where q = − |e|, where e is the elementary charge,
e2 2 ∆v 2 dω 1 2 ∆v 2 dω
dNω = ≈ (8.153)
4πε0 h̄c 3π c ω 137 3π c ω
where we used the value of the fine structure constant α = e2 /(4πε0 h̄c) ≈ 1/137.
Even if the number of photons becomes infinite when ω → 0, these photons have
negligible energies so that the total radiated energy is still finite.
E ND OF EXAMPLE 8.3
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
ϕ(t ) = ω0 t (8.154a)
x (t ) = a[ x̂1 cos ϕ(t ) + x̂2 sin ϕ(t )] (8.154b)
v(t ) = ẋ (t ) = aω0 [− x̂1 sin ϕ(t ) + x̂2 cos ϕ(t )] (8.154c)
v = |v| = aω0 (8.154d)
v̇(t ) = ẍ (t ) = −aω20 [ x̂1 cos ϕ(t ) + x̂2 sin ϕ(t )]
(8.154e)
v̇ = |v̇| = aω20 (8.154f)
where α is the angle between x − x and the normal to the plane of the particle
orbit (see Figure 8.10). From the above expressions we obtain
(x − x ) · v = x − x v sin α cos ϕ (8.156a)
(x − x ) · v̇ = − x − x v̇ sin α sin ϕ = x − x v̇ cos θ
(8.156b)
where in the last step we simply used the definition of a scalar product and the
fact that the angle between v̇ and x − x is θ.
The power flux is given by the Poynting vector, which, with the help of
Formula (8.88) on page 132, can be written
1 1 x − x
S= (E × B) = |E|2 (8.157)
µ0 cµ0 |x − x |
Inserting this into Equation (8.133) on page 142, we obtain
dU rad (α, ϕ) |x − x | s 2
= |E| (8.158)
dt cµ0
where the retarded distance s is given by expression (8.66) on page 127. With
the radiation part of the electric field, expression (8.89) on page 132, inserted,
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
x2
(t, x) x − x
x
v
q
θ (t , x )
a
v̇
α ϕ(t )
0 x1
x3
and using (8.156a) and (8.156b) on the preceding page, one finds, after some
algebra, that
2 2
2 2 1 − v sin α cos ϕ − 1 − v2 sin2 α sin2 ϕ
dU (α, ϕ) µ0 q v̇
rad c c
= 2
5 (8.159)
dt 16π c 1 − v sin α cos ϕ c
The angles θ and ϕ vary in time during the rotation, so that θ refers to a moving
coordinate system. But we can parametrise the solid angle dΩ in the angle ϕ
and the (fixed) angle α so that dΩ = sin α dα dϕ. Integration of Equation (8.159)
over this dΩ gives, after some cumbersome algebra, the angular integrated
expression
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
Cyclotron radiation
For a non-relativistic speed v c, Equation (8.159) on the preceding page
reduces to
dU rad (α, ϕ) µ0 q 2 v̇2
= (1 − sin2 α sin2 ϕ) (8.161)
dt 16π2 c
But, according to Equation (8.156b) on page 146
where θ is defined in Figure 8.10 on the preceding page. This means that we
can write
dU rad (θ) µ0 q 2 v̇2 µ0 q 2 v̇2
= (1 − cos 2
θ) = sin2 θ (8.163)
dt 16π2 c 16π2 c
Consequently, a fixed observer near the orbit plane will observe cyclotron
radiation twice per revolution in the form of two equally broad pulses of radi-
ation with alternating polarisation.
Synchrotron radiation
When the particle is relativistic, v c, the denominator in Equation (8.159)
on the previous page becomes very small if sin α cos ϕ ≈ 1, which defines the
forward direction of the particle motion (α ≈ π/2, ϕ ≈ 0). Equation (8.159) on
the preceding page then becomes
which means that an observer near the orbit plane sees a very strong pulse
followed, half an orbit period later, by a much weaker pulse.
The two cases represented by Equation (8.163) above and Equation (8.164)
are very important results since they can be used to determine the character-
istics of the particle motion both in particle accelerators and in astrophysical
objects where a direct measurement of particle velocities are impossible.
In the orbit plane (α = π/2), Equation (8.159) on the previous page gives
2 2
2 2 1 − v cos ϕ − 1 − v2 sin2 ϕ
dU (π/2, ϕ) µ0 q v̇
rad c c
= 2
5 (8.165)
dt 16π c 1 − cos ϕ
v
c
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
x2
(t, x)
x − x
∆θ
v
q
a ∆θ (t , x )
v̇
ϕ(t )
0 x1
x3
F IGURE 8.11: When the observation point is in the plane of the particle
orbit, i.e., α = π/2 the lobe width is given by ∆θ.
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
This angular interval is swept by the charge during the time interval
∆θ
∆t = (8.170)
ω0
during which the particle moves a length interval
∆θ
∆l = v∆t = v (8.171)
ω0
in the direction toward the observer who therefore measures a pulse width of
length
∆l v∆t v v ∆θ v 1
∆t = ∆t − = ∆t − = 1− ∆t = 1 − ≈ 1−
c c c c ω0 c γω0
1 − vc 1 + vc 1 v 2 1 1 1
= v ≈ 1− 2 = 3
1+ γω 0 c
) *+ , 2γω 0 2γ ω0
c
) *+ , 1/γ2
≈2
(8.172)
As a general rule, the spectral width of a pulse of length ∆t is ∆ω 1/∆t.
In the ultra-relativistic synchrotron case one can therefore expect frequency
components up to
1
ωmax ≈ = 2γ3 ω0 (8.173)
∆t
A spectral analysis of the radiation pulse will will therefore exhibit a (broadened)
line spectrum of Fourier components nω 0 from n = 1 up to n ≈ 2γ3 .
When many charged particles, N say, contribute to the radiation, we can
have three different situations depending on the relative phases of the radiation
fields from the individual particles:
1. All N radiating particles are spatially much closer to each other than a
typical wavelength. Then the relative phase differences of the individual
electric and magnetic fields radiated are negligible and the total radi-
ated fields from all individual particles will add up to become N times
that from one particle. This means that the power radiated from the N
particles will be N 2 higher than for a single charged particle. This is
called coherent radiation.
2. The charged particles are perfectly evenly distributed in the orbit. In this
case the phases of the radiation fields cause a complete cancellation of
the fields themselves. No radiation escapes.
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
Virtual photons
Let us consider a charge q moving with constant, high velocity v(t ) along the
x1 axis. According to Formula (8.102) on page 136 and Figure 8.12 on the
following page, the perpendicular component along the x 3 axis of the electric
field from this moving charge is
q v2
E⊥ = E3 = 1 − 2 (x − x0 ) · x̂3 (8.176)
4πε0 s3 c
Utilising expression (8.97a) on page 134 and simple geometrical relations, we
can rewrite this as
q b
E⊥ =
(8.177)
4πε0 γ2 (vt)2 + b2 /γ2 3/2
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
vt
q
v = v x̂1
b θ0
|x − x0 |
E⊥ x̂3
This represents a contracted field, approaching the field of a plane wave. The
passage of this field “pulse” corresponds to a frequency distribution of the field
energy. Fourier transforming, we obtain
1 ∞ q bω bω
Eω,⊥ = dt E⊥ (t) e = 2
iωt
K1 (8.178)
2π −∞ 4π ε0 bv vγ vγ
Here, K1 is the Kelvin function (Bessel function of the second kind with ima-
ginary argument) which behaves in such a way for small and large arguments
that
q
Eω,⊥ ∼ 2 , bω vγ (8.179a)
4π ε0 bv
Eω,⊥ ∼ 0, bω vγ (8.179b)
showing that the “pulse” length is of the order b/(vγ).
Due to the equipartition of the field energy into the electric and magnetic
fields, the total field energy can be written
bmax ∞
Ũ = ε0 E⊥2 d3x = ε0 E⊥2 vdt 2πb db (8.180)
V bmin −∞
where the volume integration is over the plane perpendicular to v. With the
use of Parseval’s identity for Fourier transforms, Formula (7.35) on page 107,
we can rewrite this as
∞ bmax ∞
Ũ = Ũω dω = 4πε0 v Eω,⊥ 2 dω 2πb db
0 bmin 0
2 ∞ vγ/ω (8.181)
q db
≈ 2 dω
2π ε0 v 0 bmin b
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
p2 p2
p1 p1
∇ · D = ρ(t, x) (8.188a)
∂B
∇×E = − (8.188b)
∂t
∇·B = 0 (8.188c)
∂D
∇×H = + j(t, x) (8.188d)
∂t
Assuming for simplicity that the electric permittivity ε and the magnetic
permeability µ, and hence the relative permittivity κ and the relative permeab-
ility κm all have fixed values, independent on time and space, for each type
of material we consider, we can derive the general telegrapher’s equation [cf.
Equation (2.33) on page 31]
∂2 E ∂E ∂2 E
− σµ − εµ 2 = 0 (8.189)
∂ζ 2 ∂t ∂t
∂2 E ∂2 E
− εµ =0 (8.190)
∂ζ 2 ∂t2
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
1 1 c
vϕ = √ = √ =√ (8.191)
εµ κε0 κm µ0 κκm
√
where, according to Equation (1.11) on page 6, c = 1/ ε0 µ0 is the speed of
light, i.e., the phase speed of electromagnetic waves in vacuum, then the gen-
eral solution to each component of Equation (8.190) on the preceding page
E = E0 ei(k·x−ωt) (8.195)
where now k is the wave vector in the medium given by Equation (8.194).
With these definitions, the vacuum formula for the associated magnetic field,
Equation (2.40) on page 32,
√ 1 1
B = εµ k̂ × E = k̂ × E = k × E (8.196)
vϕ ω
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
∂ω
vg = (8.198)
∂k
has a unique value for each frequency component, and is different from v ϕ .
Except in regions of anomalous dispersion, v ϕ is always smaller than c. In
a gas of free charges, such as a plasma, the refractive index is given by the
expression
ω2p
n2 (ω) = 1 − (8.199)
ω2
where
Nσ q2σ
ω2p = ∑ (8.200)
σ ε0 mσ
is the plasma frequency. Here mσ and Nσ denote the mass and number density,
respectively, of charged particle species σ. In an inhomogeneous plasma, N σ =
Nσ (x) so that the refractive index and also the phase and group velocities are
space dependent. As can be easily seen, for each given frequency, the phase
and group velocities in a plasma are different from each other. If the frequency
ω is such that it coincides with ω p at some point in the medium, then at that
point vϕ → ∞ while vg → 0 and the wave Fourier component at ω is reflected
there.
Vavilov-Čerenkov radiation
As we saw in Subsection 8.1, a charge in uniform, rectilinear motion in va-
cuum does not give rise to any radiation; see in particular Equation (8.100a)
on page 135. Let us now consider a charge in uniform, rectilinear motion in a
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
medium with electric properties which are different from those of a (classical)
vacuum. Specifically, consider a medium where
c 1
vϕ = =√ <c (8.202)
n εµ0
Hence, in this particular medium, the speed of propagation of (the phase planes
of) electromagnetic waves is less than the speed of light in vacuum, which we
know is an absolute limit for the motion of anything, including particles. A
medium of this kind has the interesting property that particles, entering into
the medium at high speeds |v|, which, of course, are below the phase speed
in vacuum, can experience that the particle speeds are higher than the phase
speed in the medium. This is the basis for the Vavilov- Čerenkov radiation that
we shall now study.
If we recall the general derivation, in the vacuum case, of the retarded (and
advanced) potentials in Chapter 3 and the Liénard-Wiechert potentials, Equa-
tions (8.65) on page 127, we realise that we obtain the latter in the medium
by a simple formal replacement c → c/n in the expression (8.66) on page 127
for s. Hence, the Liénard-Wiechert potentials in a medium characterized by a
refractive index n, are
1 q 1 q
φ(t, x) = = (8.203a)
4πε0 |x − x | − n (x−x )·v 4πε0 s
c
1 q v 1 q v
A(t, x) = = (8.203b)
4πε0 c2 |x − x | − n (x−x )·v 4πε0 c2 s
c
where now
(x − x ) · v
s = x−x −n (8.204)
c
The need for the absolute value of the expression for s is obvious in the case
when v/c ≥ 1/n because then the second term can be larger than the first term;
if v/c 1/n we recover the well-known vacuum case but with modified phase
speed. We also note that the retarded and advanced times in the medium are
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
l = v(t − t ) (8.209)
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
x(t)
θc αc q v
x (t )
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
In this manner, using jω from Equation (8.212) on the previous page, the result-
ing Fourier transforms of the Vavilov- Čerenkov magnetic and electric radiation
fields can be calculated from the expressions (7.11) and (7.22) on page 102,
respectively.
The total energy content is then obtained from Equation (7.35) on page 107
(integrated over a closed sphere at large distances). For a Fourier component
one obtains [cf. Equation (7.38) on page 108]
2
1
−ik·x 3
Uω dΩ ≈
rad
(jω × k)e d x dΩ
4πε0 nc V
2 (8.214)
q 2 nω2 ∞ ωx
= exp i − kx cos θ dx sin θ dΩ
2
16π3 ε0 c3 −∞ v
where θ is the angle between the direction of motion, x̂ 1 , and the direction to
the observer, k̂. The integral in (8.214) is singular of a “Dirac delta type.” If
we limit the spatial extent of the motion of the particle to the closed interval
[−X, X] on the x axis we can evaluate the integral to obtain
Xω
q 2 nω2 sin2 θ sin2 1 − nvc cos θ v
Uω dΩ =
rad
2 dΩ (8.215)
4π3 ε0 c3 1 − nv cos θ ω
c v
which has a maximum in the direction θc as expected. The magnitude of this
maximum grows and its width narrows as X → ∞. The integration of (8.215)
over Ω therefore picks up the main contributions from θ ≈ θ c . Consequently,
we can set sin2 θ ≈ sin2 θc and the result of the integration is
π 1
Ũωrad = 2π Uωrad (θ) sinθ dθ = cos θ = −ξ = 2π Uωrad (ξ) dξ
−1
0
1 sin2 1 + nvξ Xω (8.216)
q nω sin θc
2 2 2
c v
≈ 2 dξ
2π ε0 c
2 3
−1 ω
1 + nvξ
c v
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
leading to the final approximate result for the total energy loss in the frequency
interval (ω, ω + dω)
q 2 X c2
Ũω dω =
rad
1 − 2 2 ω dω (8.218)
2πε0 c2 n v
This result was derived under the assumption that v/c > 1/n(ω), i.e., under the
condition that the expression inside the parentheses in the right hand side is
positive. For all media it is true that n(ω) → 1 when ω → ∞, so there exist al-
ways a highest frequency for which we can obtain Vavilov- Čerenkov radiation
from a fast charge in a medium. Our derivation above for a fixed value of n is
valid for each individual Fourier component.
Bibliography
[1] H. A LFVÉN AND N. H ERLOFSON, Cosmic radiation and radio stars, Physical
Review, 78 (1950), p. 616.
[5] J. D. JACKSON, Classical Electrodynamics, third ed., John Wiley & Sons, Inc.,
New York, NY . . . , 1999, ISBN 0-471-30932-X.
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
[10] J. VANDERLINDE, Classical Electromagnetic Theory, John Wiley & Sons, Inc.,
New York, Chichester, Brisbane, Toronto, and Singapore, 1993, ISBN 0-471-
57269-1.
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
F
Formulae
∇·D = ρ (F.1)
∇·B = 0 (F.2)
∂
∇×E = − B (F.3)
∂t
∂
∇×H = j+ D (F.4)
∂t
Constitutive relations
D = εE (F.5)
B
H= (F.6)
µ
j = σE (F.7)
P = ε0 χE (F.8)
B = ∇×A (F.9)
∂
E = −∇φ − A (F.10)
∂t
163
i i
i
164 F ORMULAE
ωµ0 eik|x|
ω (x) = −
Brad pω × k (F.17)
4π |x|
1 eik|x|
ω (x) = −
Erad (pω × k) × k (F.18)
4πε0 |x|
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
µ0 eik|x|
ω (x) = −
Brad (mω × k) × k (F.19)
4π |x|
k eik|x|
ω (x) =
Erad mω × k (F.20)
4πε0 c |x|
iµ0 ω eik|x|
ω (x) =
Brad k · Qω × k (F.21)
8π |x|
i eik|x|
Erad
ω (x) = k · Qω × k × k (F.22)
8πε0 |x|
v
s = x − x − (x − x ) · (F.25)
c
v
x − x0 = (x − x ) − |x − x | (F.26)
c
∂t |x − x |
= (F.27)
∂t x s
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
166 F ORMULAE
F.3.5 Four-velocity
dx µ
uµ = = γ(c, v) (F.35)
dτ
F.3.6 Four-momentum
µ µ E
p = m0 u = ,p (F.36)
c
F.3.8 Four-potential
φ
Aµ = ,A (F.38)
c
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
where x̂i , i = 1, 2, 3 is the ith unit vector and x̂ 1 ≡ x̂, x̂2 ≡ ŷ, and x̂3 ≡ ẑ. In
component (tensor) notation ∇ can be written
∂ ∂ ∂ ∂ ∂ ∂
∇i = ∂i = , , = , , (F.41)
∂x1 ∂x2 ∂x3 ∂x ∂y ∂z
x̂1 = sin θ cos ϕr̂ + cos θ cos ϕθ̂ − sin ϕϕ̂ (F.43a)
x̂2 = sin θ sin ϕr̂ + cos θ sin ϕθ̂ + cos ϕϕ̂ (F.43b)
x̂3 = cos θr̂ − sin θθ̂ (F.43c)
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
168 F ORMULAE
Volume element
d3x = dV = drdS = r2 dr dΩ (F.47)
a · b = b · a = δi j ai b j = ab cos θ (F.48)
a × b = −b × a = i jk a j bk x̂i (F.49)
a · (b × c) = (a × b) · c (F.50)
a × (b × c) = b(a · c) − c(a · b) (F.51)
a × (b × c) + b × (c × a) + c × (a × b) = 0 (F.52)
(a × b) · (c × d) = a · [b × (c × d)] = (a · c)(b · d) − (a · d)(b · c) (F.53)
(a × b) × (c × d) = (a × b · d)c − (a × b · c)d (F.54)
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
Special identities
In the following x = xi x̂i and x = xi x̂i are radius vectors, k an arbitrary constant
vector, a = a(x) an arbitrary vector field, ∇ ≡ ∂x∂ i x̂i , and ∇ ≡ ∂x∂ x̂i .
i
∇·x = 3 (F.65)
∇×x = 0 (F.66)
∇(k · x) = k (F.67)
x
∇|x| = (F.68)
|x|
x − x
∇ |x − x | =
= −∇ |x − x | (F.69)
|x − x |
1 x
∇ =− 3 (F.70)
|x| |x|
1 x − x 1
∇ =− = −∇ (F.71)
|x − x | |x − x |3 |x − x |
x 1
∇· = −∇2 = 4πδ(x) (F.72)
|x|3 |x|
x − x 1
∇· = −∇ 2
= 4πδ(x − x ) (F.73)
|x − x |3 |x − x |
k 1 k·x
∇ = k· ∇ =− 3 (F.74)
|x| |x| |x|
x k·x
∇× k× = −∇ if |x| = 0 (F.75)
|x|3 |x|3
k 1
∇ 2
= k∇ 2
= −4πkδ(x) (F.76)
|x| |x|
∇ × (k × a) = k(∇ · a) + k × (∇ × a) − ∇(k · a) (F.77)
Integral relations
Let V(S ) be the volume bounded by the closed surface S (V). Denote the 3-
dimensional volume element by d3x(≡ dV) and the surface element, directed
along the outward pointing surface normal unit vector n̂, by dS(≡ d 2x n̂). Then
(∇ · a) d x =
3
dS · a (F.78)
V S
(∇α) d3x = dS α (F.79)
V S
(∇ × a) d3x = dS × a (F.80)
V S
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
170 F ORMULAE
If S (C) is an open surface bounded by the contour C(S ), whose line ele-
ment is dl, then
α dl = dS × ∇α (F.81)
C S
a · dl = dS · (∇ × a) (F.82)
C S
Bibliography
[1] G. B. A RFKEN AND H. J. W EBER, Mathematical Methods for Physicists, fourth,
international ed., Academic Press, Inc., San Diego, CA . . . , 1995, ISBN 0-12-
059816-7.
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
M
Mathematical
Methods
M.1.1 Vectors
Radius vector
A vector can be represented mathematically in a number of different ways.
One suitable representation is in terms of an ordered N-tuple, or row vector, of
the coordinates x N where N is the dimensionality of the space under consider-
ation. The most basic vector is the radius vector which is the vector from the
origin to the point of interest. Its N-tuple representation simply enumerates the
coordinates which describe this point. In this sense, the radius vector from the
origin to a point is synonymous with the coordinates of the point itself.
171
i i
i
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
M.1.2 Fields
A field is a physical entity which depends on one or more continuous paramet-
ers. Such a parameter can be viewed as a “continuous index” which enumer-
ates the “coordinates” of the field. In particular, in a field which depends on
the usual radius vector x of R 3 , each point in this space can be considered as
one degree of freedom so that a field is a representation of a physical entity
which has an infinite number of degrees of freedom.
Scalar fields
We denote an arbitrary scalar field in R 3 by
def
α(x) = α(x1 , x2 , x3 ) ≡ α(xi ) (M.6)
def
α(x0 , x1 , x2 , x3 ) ≡ α(xµ ) (M.7)
which indicates that the four-scalar α depends on all four coordinates spanning
this space. Since a four-scalar has the same value at a given point regardless
of coordinate system, it is also called an invariant.
Analogous to the transformation rule, Equation (M.5) on the preceding
page, for the differential dx µ , the transformation rule for the differential oper-
ator ∂/∂xµ under a transformation x µ → xµ becomes
∂ ∂xν ∂
= (M.8)
∂xµ ∂xµ ∂xν
which, again, follows trivially from the rules of differentiation.
Vector fields
We can represent an arbitrary vector field a(x) in R 3 as follows:
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
∂xµ ν
yµ = y (M.13)
∂xν
i.e., in the same way as the differential coordinate element dx µ transforms
according to Equation (M.5) on page 172.
The analogous requirement for a covariant four-vector is that it transforms,
during the change from Σ to Σ , according to the rule
∂xν
yµ = yν (M.14)
∂xµ
i.e., in the same way as the differential operator ∂/∂x µ transforms according to
Equation (M.8) on the previous page.
Tensor fields
We denote an arbitrary tensor field in R 3 by A(x). This tensor field can be
represented in a number of ways, for instance in the following matrix form:
⎛ ⎞
def A11 (x) A12 (x) A13 (x)
Ai j (xk ) ≡ ⎝A21 (x) A22 (x) A23 (x)⎠ (M.15)
A31 (x) A32 (x) A33 (x)
Strictly speaking, the tensor field described here is a tensor of rank two.
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
In fact, tensors may have any rank n. In this picture a scalar is considered
to be a tensor of rank n = 0 and a vector a tensor of rank n = 1. Consequently,
the notation where a vector (tensor) is represented in its component form is
called the tensor notation. A tensor of rank n = 2 may be represented by a two-
dimensional array or matrix whereas higher rank tensors are best represented
in their component forms (tensor notation).
T− n̂ = −T n̂ (M.20)
Using (M.20) and Newton’s second law, we find that the matter of mass m, which at a
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
x3
n̂
d2x
x2
x1
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
form as follows
3
T n̂j = (T n̂) j = ∑ ni T i j ≡ ni T i j (M.25)
i=1
Using Equation (M.25) above, we find that the component of the vector T n̂ in the
direction of an arbitrary unit vector m̂ is
T n̂m̂ = T n̂ · m̂
1 2
3 3 3 (M.26)
= ∑ T n̂j m j = ∑ ∑ ni Ti j m j ≡ ni T i j m j = n̂· T · m̂
j=1 j=1 i=1
Hence, the jth component of the vector T x̂i , here denoted T i j , can be interpreted as the
i jth component of a tensor T. Note that T n̂m̂ is independent of the particular coordinate
system used in the derivation.
We shall now show how one can use the momentum law (force equation) to derive
the equation of motion for an arbitrary element of mass in the body. To this end we
consider a part V of the body. If the external force density (force per unit volume) is
denoted by f and the velocity for a mass element dm is denoted by v, we obtain
d
v dm = f d3x + T n̂ d2x (M.27)
dt V V S
where, in the last step, Equation (M.25) was used. Setting dm = ρ d3x and using the
divergence theorem on the last term, we can rewrite the result as
d ∂T i j 3
ρ v j d3x = f j d3x + dx (M.29)
V dt V V ∂xi
Since this formula is valid for any arbitrary volume, we must require that
d ∂T i j
ρ vj − fj − =0 (M.30)
dt ∂xi
or, equivalently
∂v j ∂T i j
ρ + ρv · ∇v j − f j − =0 (M.31)
∂t ∂xi
Note that ∂v j /∂t is the rate of change with time of the velocity component v j at a fixed
point x = (x1 , x1 , x3 ).
E ND OF EXAMPLE M.1
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
def
aµ (xκ ) ≡ gµν aν (xκ ) (M.32)
This rule is often called lowering of index. The raising of index analogue of
the index lowering rule is:
def
aµ (xκ ) ≡ gµν aν (xκ ) (M.33)
More generally, the following lowering and raising rules hold for arbitrary
rank n mixed tensor fields:
gµk νk Aνν1k+1
ν2 ...νk−1 νk κ ν1 ν2 ...νk−1 κ
νk+2 ...νn (x ) = Aµk νk+1 ...νn (x ) (M.34)
Successive lowering and raising of more than one index is achieved by a re-
peated application of this rule. For example, a dual application of the lowering
operation on a rank 2 tensor in contravariant form yields
i.e., the same rank 2 tensor in covariant form. This operation is also known as
a tensor contraction.
The 4D Lorentz space L4 has a simple metric which can be described either by the
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
metric tensor
⎧
⎪
⎨1 if µ = ν = 0
gµν = −1 if µ = ν = i = j = 1, 2, 3 (M.37)
⎪
⎩
0 if µ =
ν
which, in matrix notation, is represented as
⎛ ⎞
1 0 0 0
⎜0 −1 0 0⎟
(gµν ) = ⎜
⎝0 0 −1 0 ⎠
⎟ (M.38)
0 0 0 −1
i.e., a matrix with a main diagonal that has the sign sequence, or signature, {+, −, −, −}
or
⎧
⎪
⎨−1 if µ = ν = 0
gµν = 1 if µ = ν = i = j = 1, 2, 3 (M.39)
⎪
⎩
0 if µ = ν
which, in matrix notation, is represented as
⎛ ⎞
−1 0 0 0
⎜ 0 1 0 0⎟
(gµν ) = ⎜
⎝ 0 0 1 0⎠
⎟ (M.40)
0 0 0 1
i.e., a matrix with signature {−, +, +, +}.
Consider an arbitrary contravariant four-vector aν in this space. In component form it
can be written:
def
aν ≡ (a0 , a1 , a2 , a3 ) = (a0 , a) (M.41)
According to the index lowering rule, Equation (M.32) on the preceding page, we
obtain the covariant version of this vector as
def
aµ ≡ (a0 , a1 , a2 , a3 ) = gµν aν (M.42)
In the {+, −, −, −} metric we obtain
µ=0: a0 = 1 · a0 + 0 · a1 + 0 · a2 + 0 · a3 = a0 (M.43)
µ=1: a1 = 0 · a − 1 · a + 0 · a + 0 · a = −a
0 1 2 3 1
(M.44)
µ=2: a2 = 0 · a + 0 · a − 1 · a + 0 · a = −a
0 1 2 3 2
(M.45)
µ=3: a3 = 0 · a + 0 · a + 0 · a + 1 · a = −a
0 1 2 3 3
(M.46)
or
aµ = (a0 , a1 , a2 , a3 ) = (a0 , −a1 , −a2 , −a3 ) = (a0 , −a) (M.47)
Radius 4-vector itself in L4 and in this metric is given by
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
E ND OF EXAMPLE M.2
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
def def
A ≡ aR + iaI = aR âR + iaI âI ≡ A Â ∈ C3 (M.55)
def def
A2 ≡ A · A = a2R − a2I + 2iaR · aI ≡ A2 ∈ C (M.56)
Using this in Equation (M.55) above, we see that we can interpret this so that the
complex unit vector is
A aR aI
 = =( âR + i ( âI
A a2R − a2I + 2iaR · aI a2R − a2I + 2iaR · aI
( (
aR a2R − a2I − 2iaR · aI aI a2R − a2I − 2iaR · aI
= âR + i âI ∈ C3
a2R + a2I a2R + a2I
(M.58)
On the other hand, the definition of the scalar product in terms of the inner product of
complex vector with its own complex conjugate yields
def
|A|2 ≡ A · A∗ = a2R + a2I = |A|2 (M.59)
A aR aI
 = =( âR + i ( âI
|A| a2R + a2I a2R + a2I
( ( (M.60)
aR a2R + a2I aI a2R + a2I
= âR + i âI ∈ C3
a2R + a2I a2R + a2I
E ND OF EXAMPLE M.3
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
In L4 the metric tensor attains a simple form [see Example M.2 on page 179] and,
hence, the scalar product in Equation (M.53) on page 180 can be evaluated almost
trivially. For the {+, −, −, −} signature it becomes
The important scalar product of the L4 radius four-vector with itself becomes
which is the indefinite, real norm of L4 . The L4 metric is the quadratic differential
form
E ND OF EXAMPLE M.4
where κ = κ(ct, r, θ, φ) and λ = λ(ct, r, θ, φ). In such a space, the metric takes the form
In general relativity the metric tensor is not given a priori but is determined by the
Einstein equations.
E ND OF EXAMPLE M.5
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
Dyadic product
The dyadic product field A(x) ≡ a(x)b(x) with two juxtaposed vector fields
a(x) and b(x) is the outer product of a and b. Operating on this dyad from the
right and from the left with an inner product of an vector c one obtains
def def
A · c ≡ ab · c ≡ a(b · c) (M.66a)
def def
c · A ≡ c · ab ≡ (c · a)b (M.66b)
which means that we can represent the tensor A(x) in matrix form as
⎛ ⎞
a 1 b 1 a 1 b 2 a 1 b 3
Ai j (xk ) = ⎝a1 b2 a2 b2 a2 b3 ⎠ (M.68)
a1 b3 a3 b2 a3 b3
which we identify with expression (M.15) on page 174, viz. a tensor in matrix
notation.
Vector product
The vector product or cross product of two arbitrary 3D vectors a and b in
ordinary R3 space is the vector
c = a × b = i jk a j bk x̂i (M.69)
a · b = ab sin θ ê (M.70)
where θ is the angle between a and b and ê is a unit vector perpendicular to the
plane spanned by a and b.
A spatial reversal of the coordinate system (x 1 , x2 , x3 ) = (−x1 , −x2 , −x3 )
changes sign of the components of the vectors a and b so that in the new
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
def ∂ def
∇ ≡ x̂i ≡ ∂ (M.71)
∂xi
where x̂i is the ith unit vector in a Cartesian coordinate system. Since the
operator in itself has vectorial properties, we denote it with a boldface nabla.
In “component” notation we can write
∂ ∂ ∂
∂i = , , (M.72)
∂x1 ∂x2 ∂x3
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
and
yµ = ∂µ xν yν (M.76)
respectively.
With the help of the del operator we can define the gradient, divergence
and curl of a tensor (in the generalised sense).
The gradient
The gradient of an R3 scalar field α(x), denoted ∇α(x), is an R 3 vector field
a(x):
From this we see that the boldface notation for the nabla and del operators is
very handy as it elucidates the 3D vectorial property of the gradient.
In 4D, the four-gradient is a covariant vector, formed as a derivative of a
four-scalar field α(x µ ), with the following component form:
∂α(xν )
∂µ α(xν ) = (M.81)
∂xµ
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
∂
∇ = x̂i = ∂ (M.82)
∂xi
Using this, the “unprimed” version, Equation (M.71) on page 184, and elementary
rules of differentiation, we obtain the following two very useful results:
∂|x − x | x − x ∂|x − x |
∇ |x − x | = x̂i = = − x̂i
∂xi
|x − x | ∂xi (M.83)
= −∇ |x − x |
and
1 x − x 1
∇ =− = −∇ (M.84)
|x − x | |x − x |3 |x − x |
E ND OF EXAMPLE M.7
The divergence
We define the 3D divergence of a vector field in R 3 as
∂ai (x)
∇ · a(x) = ∂ · x̂ j a j (x) = δi j ∂i a j (x) = ∂i ai (x) = = α(x) (M.85)
∂xi
∂aµ (xν )
∂µ aµ (xν ) = ∂µ aµ (xν ) = (M.86)
∂xµ
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
The Laplacian
The 3D Laplace operator or Laplacian can be described as the divergence of
the gradient operator:
∂ ∂ ∂2 3
∂2
∇2 = ∆ = ∇ · ∇ = x̂i · x̂ j = δi j ∂i ∂ j = ∂2i = 2 ≡ ∑ 2 (M.88)
∂xi ∂x j ∂xi i=1 ∂xi
The symbol ∇2 is sometimes read del squared. If, for a scalar field α(x),
∇2 α < 0 at some point in 3D space, it is a sign of concentration of α at that
point.
The curl
In R3 the curl of a vector field a(x), denoted ∇ × a(x), is another R 3 vector field
b(x) which can be defined in the following way:
∂ak (x)
∇ × a(x) = i jk x̂i ∂ j ak (x) = i jk x̂i = b(x) (M.90)
∂x j
where use was made of the Levi-Civita tensor, introduced in Equation (M.18)
on page 175.
The covariant 4D generalisation of the curl of a four-vector field a µ (xν ) is
the antisymmetric four-tensor field
Gµν (xκ ) = ∂µ aν (xκ ) − ∂ν aµ (xκ ) = −Gνµ (xκ ) (M.91)
A vector with vanishing curl is said to be irrotational.
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
Using the definition of the R3 curl, Equation (M.90) on the preceding page, and the
gradient, Equation (M.80) on page 185, we see that
∇ × [∇α(x)] ≡ 0 (M.94)
E ND OF EXAMPLE M.10
With the use of the definitions of the divergence (M.85) and the curl, Equation (M.90)
on the preceding page, we find that
Using the definition for the Levi-Civita symbol, defined by Equation (M.18) on
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
E ND OF EXAMPLE M.11
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
After differentiating the left and right hand sides of this definition and setting
them equal we obtain
∂H ∂H ∂H ∂L ∂L ∂L
dpi + dqi + dt = q̇i dpi + pi dq̇i − dqi − dq̇i − dt (M.105)
∂pi ∂qi ∂t ∂qi ∂q̇i ∂t
According to the definition of p i , Equation (M.102) above, the second and
fourth terms on the right hand side cancel. Furthermore, noting that according
to Equation (M.103) the third term on the right hand side of Equation (M.105)
above is equal to − ṗi dqi and identifying terms, we obtain the Hamilton equa-
tions:
∂H dqi
= q̇i = (M.106a)
∂pi dt
∂H dpi
= − ṗi = − (M.106b)
∂qi dt
Bibliography
[1] G. B. A RFKEN AND H. J. W EBER, Mathematical Methods for Physicists, fourth,
international ed., Academic Press, Inc., San Diego, CA . . . , 1995, ISBN 0-12-
059816-7.
[2] R. A. D EAN, Elements of Abstract Algebra, John Wiley & Sons, Inc.,
New York, NY . . . , 1967, ISBN 0-471-20452-8.
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
[5] B. S PAIN, Tensor Calculus, third ed., Oliver and Boyd, Ltd., Edinburgh and Lon-
don, 1965, ISBN 05-001331-9.
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
Index
193
i i
i
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
195
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i
i
Downloaded from http://www.plasma.uu.se/CED/Book Draft version released 23rd January 2003 at 18:57.
i i
i
197
Draft version released 23rd January 2003 at 18:57. Downloaded from http://www.plasma.uu.se/CED/Book
i i