Q4-Another Form
Q4-Another Form
Ph501
Electrodynamics
Problem Set 8
Kirk T. McDonald
(2001)
kirkmcd@princeton.edu
http://kirkmcd.princeton.edu/examples/
Princeton University 2001 Ph501 Set 8, Problem 1 1
Deduce the time-dependence of the electric and magnetic fields, E and B, observed at
a point (r, θ = 0, z = 0) in a cylindrical coordinate system about the wire. Use your
expressions to discuss the fields in the two limiting cases that ct r and ct = r + ,
where c is the speed of light and r.
The related, but more intricate case of a solenoid with a linearly rising current is
considered in http://kirkmcd.princeton.edu/examples/solenoid.pdf
Princeton University 2001 Ph501 Set 8, Problem 2 2
J = ṗ. (2)
The radiation fields will be deduced from the retarded vector potential,
1 [J] 1 [ṗ]
A= dVol = dVol, (3)
c r c r
which is a solution of the (Lorenz gauge) wave equation,
1 ∂ 2A 4π
∇2 A − 2 2
= − J. (4)
c ∂t c
Suppose that the Hertzian electric dipole vector p has oscillatory time dependence,
to show that
ei(kR−ωt) 1
A = −iω pω (r) 1 + r · n̂ − ik + ... dVol , (7)
cR R
where no assumption is made that R source size or that R λ = 2π/k = 2πc/ω.
Consider now only the leading term in this expansion, which corresponds to electric
dipole radiation. Introducing the total electric dipole moment,
P≡ pω (r ) dVol , (8)
1
Some consideration of the related topics of Hertz vectors and scalars is given in the Appendix of
http://kirkmcd.princeton.edu/examples/smallloop.pdf
Princeton University 2001 Ph501 Set 8, Problem 2 3
use,
1 ∂E
B=∇×A and ∇×B= (9)
c ∂t
to show that for an observer in vacuum the electric dipole radiation fields are,,
ei(kR−ωt) i
B = k2 1+ n̂ × P, (10)
R kR
ei(kR−ωt) 1 i
E = k2 n̂ × (P × n̂) + [3(n̂ · P)n̂ − P] 2 2 − . (11)
R k R kR
Alternatively, deduce the electric field from both the scalar and vector potentials via,
1 ∂A
E = −∇φ − , (12)
c ∂t
in both the Lorenz and Coulomb gauges.
For large R,
ei(kR−ωt)
Bfar ≈ k 2 n̂ × P, Efar ≈ Bfar × n̂, (13)
R
while for small R,
Calculate the radiation fields and the radiated power according to an observer at angle
θ to the z axis in the x-z plane.
Define n̂ towards the observer, so that n̂ · ẑ = cos θ, and let l̂ = ŷ × n̂.
Show that,
ei(kr−ωt) ei(kr−ωt)
Brad = p0 k 2 (cos θ ŷ − i l̂), Erad = p0 k 2 (cos θ l̂ + i ŷ), (16)
r r
where r is the distance from the center of the dipole to the observer.
Note that for an observer in the x-y plane (n̂ = x̂), the radiation is linearly polarized,
while for an observer along the z axis it is circularly polarized.
Show that the (time-averaged) radiated power is given by,
d P c 2 4 2cp20 k 4 2p2 ω 4
= p k (1 + cos2 θ), P = = 03 . (17)
dΩ 8π 0 3 3c
This example gives another simple picture of how radiation fields are generated. The
field lines emanating from the dipole become twisted into spirals as the dipole rotates.
At large distances, the field lines are transverse...
Princeton University 2001 Ph501 Set 8, Problem 4 5
4. Magnetars
The x-ray pulsar SGR1806-20 has recently
been observed to have a period T of 7.5 s
and a relatively large “spindown” rate Ṫ = 8 × 10−11 . See, C. Kouveliotou et al., An
X-ray pulsar with a superstrong magnetic field in the soft γ-ray repeater SGR1806-20,
Nature 393, 235-237 (1998).2
Calculate the maximum magnetic field at the surface of this pulsar, assuming it to be
a standard neutron star of mass 1.4M = 2.8 × 1030 kg and radius 10 km, that the
mass density is uniform, that the spindown is due to electromagnetic radiation, and
that the angular velocity vector is perpendicular to the magnetic dipole moment of the
pulsar.
Compare the surface magnetic field strength to the so-called QED critical field strength
m2c3 /eh̄ = 4.4 × 1013 gauss, at which electron-positron pair creation processes become
highly probable.
2
http://kirkmcd.princeton.edu/examples/EM/kouveliotou_nature_393_235_98.pdf
Princeton University 2001 Ph501 Set 8, Problem 5 6
Show that for oscillatory sources, the time-average angular momentum radiated into
unit solid angle per second is (the real part of),
d L 1 3
= r [E(n̂ · B ) − B(n̂ · E)]. (20)
dt dΩ 8π
Thus, the radiated angular momentum is zero for purely transverse fields.
In eq. (11) of Prob. 2 above, we found that for electric dipole radiation there is a term
in E with E · n̂ ∝ 1/r2 . Show that for radiation by an oscillating electric dipole p,
d L ik 3
= (n̂ · p)(n̂ × p ). (21)
dt dΩ 4π
If the dipole moment p is real, eq. (21) tells us that no angular momentum is radiated.
However, when p is real, the radiation is linearly polarized and we expect it to carry
no angular momentum.
Rather, we need circular (or elliptical) polarization to have radiated angular momen-
tum.
The radiation fields (16) of Prob. 2 are elliptically polarized. Show that in this case
the radiated angular momentum distribution is,
d L k3 d L P
= − p20 sin θ l̂, and = ẑ . (22)
dt dΩ 4π dt ω
[These relations carry over into the quantum realm where a single (left-hand) circularly
polarized photon has U = h̄ω, p = h̄k, and L = h̄.]
For a another view of electromagnetic waves that carry angular momentum, see
http://kirkmcd.princeton.edu/examples/oblate_wave.pdf
Princeton University 2001 Ph501 Set 8, Problem 6 7
Show that for an observer in the x-z plane at distance r from the origin,
sin(2kr − 2ωt)
Erad = −4k 3 a2e sin θ cos θ l̂, (23)
r
The time-varying electric-dipole moment of this charge distribution with respect to the
origin has magnitude p = ae, so from Larmor’s formula (prob. 2) we know that the
(time-averaged) power in electric dipole radiation is,
2a2e2 ω 4
PE1 = . (25)
3c3
This charge distribution also has a magnetic dipole moment and an electric-quadrupole
moment (plus higher moments as well!). Calculate the total radiation fields due to the
E1, M1 and E2 moments, as well as the angular distribution of the radiated power and
the total radiated power from these three moments. In this pedagogic problem you
may ignore the interference between the various moments.
Show, for example, that the part of the radiation due only to the electric-quadrupole
moment obeys,
d PE2 a4 e2 ω 6 8a4e2 ω 6
= 5
(1 − cos4 θ), PE2 = . (26)
dΩ 2πc 5c5
Thus,
PE2 12a2 ω 2 v2
= ∝ . (27)
PE1 5c2 c2
Princeton University 2001 Ph501 Set 8, Problem 8 9
Consider the Earth-Sun system. The motion of the Earth around the Sun causes a
quadrupole moment, so gravitational radiation is emitted (although, of course, there
is no dipole gravitational radiation since the dipole moment of any system of masses
about its center of mass is zero). Estimate the time for the Earth to fall into the Sun
due to gravitational-radiation energy loss.
What is the analog of the factor ea2 that appears in the electrical-quadrupole moment
(Prob. 5) for masses m1 and m2 that are in circular motion about each other, separated
by distance a?
Also note that in Gaussian units the electrical coupling constant k in the force law
F = ke1e2/r2 has been set to 1, but for gravity k = G, Newton’s constant.
The general-relativity expression for quadrupole radiation in the present example is,3
32 G m21m22
PG2 = a4 ω 6 . (32)
5 c5 (m1 + m2)2
The extra factor of 4 compared to E2 radiation arises because the source term in the
gravitational wave equation has a factor of 16π, rather than 4π as for E&M.
3
P.C. Peters and J. Mathews, Phys. Rev. 131, 435 (1963),
http://kirkmcd.princeton.edu/examples/GR/peters_pr_131_435_63.pdf
Princeton University 2001 Ph501 Set 8, Problem 9 10
1 [J] 1 1
A(r, t) = dVol ≈ [J] dVol = J(r, t = t − r/c) dVol , (33)
c r cR cR
where R is the (large) distance from the observer to the center of the ring of radius
a. For uniform circular motion of N electrons with angular frequency ω, the current
density J is a periodic function with period T = 2π/ω, so a Fourier analysis can be
made where,
∞
J(r, t) = Jm (r)e−imωt , (34)
m=−∞
with, T
1
Jm (r) = J(r, t)eimωt dt . (35)
T 0
Then,
A(r, t) = Am (r)e−imωt , (36)
m
that depends on the square of the amplitude. Transcribing the argument there to the
present case, a Fourier analysis of the average power radiated during one period T can
be given as,
d P 1 T dP cR2 T cR2 T
= dt = |B|2 dt = B Bm e−imωt dt
dΩ T 0 dΩ 4πT 0 4πT 0 m
∞
cR2 1 T −imωt cR2
= Bm B e dt = Bm Bm
4π m T 0 4π m=−∞
∞ ∞
cR2 dPm
= |Bm |2 ≡ . (38)
2π m=0 m=0 dΩ
That is, the Fourier components of the time-averaged radiated power can be written
as,
dPm cR2 cR2 cR2
= |Bm |2 = |∇ × Am |2 = |imkn̂ × Am |2 , (39)
dΩ 2π 2π 2π
where k = ω/c and n̂ points from the center of the ring to the observer.
Evaluate the Fourier components of the vector potential and of the radiated power
first for a single electron, with geometry as in Prob. 5, and then for N electrons evenly
spaced around the ring. It will come as no surprise that a 3-dimensional problem with
charges distributed on a ring leads to Bessel functions, and we must be aware of the
integral representation,
im 2π imφ−iz cos φ
Jm (z) = e dφ. (40)
2π 0
Use the asymptotic expansion for large index and small argument,
(ex/2)m
Jm (mx) ≈ √ (m 1, x 1), (41)
2πm
to verify the suppression of the radiation for large N.
This problem was first posed (and solved via series expansions without explicit mention
of Bessel functions) by J.J. Thomson, Phil. Mag. 45, 673 (1903).4 He knew that atoms
(in what we now call their ground state) don’t radiate, and used this calculation to
support his model that the electric charge in an atom must be smoothly distributed.
This was a classical precursor to the view of a continuous probability distribution for
the electron’s position in an atom.
Thomson’s work was followed shortly by an extensive treatise by G.A. Schott, Electro-
magnetic Radiation (Cambridge U.P., 1912),5 that included analyses in term of Bessel
functions correct for any value of v/c.
These pioneering works were largely forgotten during the following era of nonrelativis-
tic quantum mechanics, and were reinvented around 1945 when interest emerged in
relativistic particle accelerators. See Arzimovitch and Pomeranchuk,6 and Schwinger.7
4
http://kirkmcd.princeton.edu/examples/EM/thomson_pm_45_673_03.pdf
5
http://kirkmcd.princeton.edu/examples/EM/schott_radiation_12.pdf
6
http://kirkmcd.princeton.edu/examples/EM/arzimovitch_jpussr_9_267_45.pdf
7
http://kirkmcd.princeton.edu/accel/schwinger.pdf
Princeton University 2001 Ph501 Set 8, Problem 10 12
14. a) Consider a full-wave “end-fire” antenna whose current distribution (along the z axis)
is,
2πz −iωt
I(z) = I0 sin e , (−L/2 < z < L/2), (51)
L
where L = λ = 2πc/ω.
Use the result of p. 182, Lecture 15 of the Notes to calculate the radiated power
“exactly”. Note that the real part of the integral vanishes, so you must evaluate the
imaginary part. Show that,
dP I 2 sin2 (π cos θ)
= 0 . (52)
dΩ 2πc sin2 θ
Sketch the radiation pattern.
Use tricks like,
1 1 1
= + (53)
1−u 2 1+u 1−u
to show that the total radiated power is,
dP
P = (54)
dΩ
(c.f., Abramowitz and Stegun, pp. 231, 244.)
http://kirkmcd.princeton.edu/examples/EM/abramowitz_and_stegun.pdf
b) Calculate the lowest-order nonvanishing multipole radiation. You may need the fact
that,
z 2 cos z dz = (z 2 − 2) sin z + 2z cos z. (55)
Show that to this order,
8π2 I02 I2
P = = 5.26 0 . (56)
15 2c 2c
which gives a sense of the accuracy of the multipole expansion.
Princeton University 2001 Ph501 Set 8, Problem 15 17
Consider an observer in the x-z plane to distinguish between the cases of electric
polarization parallel and perpendicular to the scattering plane to show that,
2 2
dσ 1 dσ⊥ cos θ
= a6 k 4 − cos θ , = a6 k 4 1 − . (59)
dΩ 2 dΩ 2
Solutions
1. The suggested approach is to calculate the retarded potentials and then take derivatives
to find the fields. The retarded scalar and vector potentials φ and A are given by,
ρ(x, t − R/c) 3 1 J(x, t − R/c) 3
φ(x, t) = d x, and A(x, t) = d x , (63)
R c R
where ρ and J are the charge and current densities, respectively, and R = |x − x|.
In the present case, we assume the wire remains neutral when the current flows (com-
pare Prob. 3, Set 4). Then the scalar potential vanishes. For the vector potential, we
see that only the component Az will be nonzero. Also, J d3 x can be rewritten as I dz for
current in a wire along√the z-axis. For an observer at (r, 0, 0) and a current element at
(0, 0, z), we have R = r2 + z 2 . Further, the condition that I is nonzero only for time
t > 0 implies that it contributes to the fields only for z such that (ct)2 > R2 = r2 + z 2.
That is, we need to evaluate the integral only for,
|z| < z0 ≡ (ct)2 − r2 , (64)
so,
√
2α 2 2α r + + 2r 2α 2
Bφ ≈ 2 , and Ez ≈ − 2 ln ≈− 2 = −Bφ . (70)
c r c r c r
In this regime, the fields have the character of radiation, with E and B of equal
magnitude, mutually orthogonal, and both orthogonal to the line of sight to the closest
point on the wire. (Because of the cylindrical geometry the radiation fields do not have
1/r dependence – which holds instead for static fields.)
In sum, the fields build up from zero only after time ct = r. The initial fields propagate
outwards at the speed of light and have the character of cylindrical waves. But at a
fixed r, the electric field dies out with time, and the magnetic field approaches the
instantaneous magnetostatic field due to the current in the wire.
Of possible amusement is a direct calculation of the vector potential for the case of a
constant current I0.
First, from Ampère’s law we know that Bφ = 2I0/cr = −∂Az /∂r, so we have that,
2I0
Az = − ln r + const. (71)
c
Whereas, if we use the integral form for the vector potential we have,
∞ ∞
1 I dz 2I0 dz
Az (r, 0, 0) = √ 0 = √
c −∞ 2
r +z 2 c 0 r + z2
2
2I0 √
= − ln r + lim ln(z + z 2 + r2 ). (72)
c z→∞
Only by ignoring the last term, which does not depend on r for a long wire, do we
recover the “elementary” result.
Princeton University 2001 Ph501 Set 8, Solution 2 20
2. The expansion,
r = R − r · n̂ + ... (73)
implies that the retarded time derivative of the polarization vector is,
[ṗ] = ṗ(r, t = t − r/c) ≈ −iωpω (r) e−iω(t−R/c+r ·n̂/c) = −iω ei(kR−ωt) pω (r ) e−ikr ·n̂
≈ −iω ei(kR−ωt)pω (r )(1 − ikr · n̂), (74)
The electric-dipole (E1) approximation is to keep only the first term of eq. (76),
We could also deduce the electric field from the general relation,
1 ∂A
E = −∇V − = −∇V + ikA. (82)
c ∂t
For this, we need to know the scalar potential V (L) , which we can deduce from the
Lorenz gauge condition:
1 ∂V (L)
∇ · A(L) + = 0. (83)
c ∂t
For an oscillatory source this becomes,
i
V (L) = − ∇ · A(L) . (84)
k
In the electric-dipole approximation (77) this yields,8
1 ik
= ei(kR−ωt)
(L)
VE1 2
− (P · n̂) (Lorenz gauge). (85)
R R
(L) P · n̂ −iωt
VE1,near ≈ e . (86)
R2
The electric field is given by,
(L) (L)
EE1 = −∇VE1 + ikAE1
i(kR−ωt)
ik 1 ik 1 2e
= (P · R)∇ei(kR−ωt) − + e i(kR−ωt)
− ∇(P · R) + k P
R2 R3 R2 R3 R
i(kR−ωt) k2 ik 1 i(kR−ωt) ik 1
= e − −3 − (P · n̂)n̂ + e − P
R R2 R3 R2 R3
ei(kR−ωt)
+k 2 P
R
i(kR−ωt)
2e i(kR−ωt) ik 1
= k n̂ × (P × n̂) + e − [P − 3(P · n̂)n̂], (87)
R R2 R3
as before. The angular distribution in the far field (for which the radial dependence is
1/R) is n̂× (P× n̂) = P− (P· n̂)n̂. The isotropic term P is due to the vector potential,
while the variable term −(P · n̂)n̂ is due to the scalar potential and is purely radial.
Spherical waves associated with a scalar potential must be radial (longitudinal), but
the transverse character of electromagnetic waves in the far field does not imply the
absence of a contribution of the scalar potential; the latter is needed (in the Lorenz
gauge) to cancel to radial component of the waves from the vector potential.
8
Equation (85) is not simply the electrostatic-dipole potential times a spherical wave because the retarded
positions at time t of the two charges of a point dipole correspond
to two different retarded times t .
(L)
For a calculation of the retarded scalar potential via V = Vol [ρ]/r, see sec. 11.1.2 of Introduction to
Electrodynamics by D.J. Griffiths.
Princeton University 2001 Ph501 Set 8, Solution 2 22
We could also work in the Coulomb gauge, meaning that we set ∇ · A(C) = 0. Recall
(Lecture 15, p. 174) that the “wave” equations for the potentials in the Coulomb gauge
are,
(C) P · n̂ −iωt
VE1 = e (Coulomb gauge). (90)
R2
One way to deduce the Coulomb-gauge vector potential is via eq. (82),
(C) i i (C)
AE1 = − EE1 − ∇VE1
k k
ei(kR−ωt) ei(kR−ωt) i(ei(kR−ωt) − e−iωt )
= −ik n̂ × (P × n̂) + + [P − 3(P · n̂)n̂]
R R2 kR3
(C)
≡ Afar + A(C)
near (Coulomb gauge). (91)
We learn that the far-zone, Coulomb gauge vector potential (i.e., the part of the vector
potential that varies as 1/R) is purely transverse, and can be written as
(C) ei(kR−ωt)
Afar = −ik n̂ × (P × n̂) (Coulomb gauge). (92)
R
Because the radiation part of the Coulomb-gauge vector potential is transverse, the
Coulomb gauge is sometimes called the “transverse” gauge.
The Coulomb-gauge scalar potential is negligible in the far zone, and we can say that
the radiation fields are entirely due to the far-zone, Coulomb-gauge vector potential.
That is,
(C) ei(kR−ωt)
Efar = ikAfar = k 2 n̂ × (P × n̂), (93)
R
i(kR−ωt)
(C) (C) 2e
Bfar = ∇ × Afar = ik × Afar = k n̂ × P. (94)
R
It is possible to choose gauges for the electromagnetic potentials such that some of
their components appear to propagate at any velocity v, as discussed by J.D. Jackson,
Am. J. Phys. 70, 917 (2002) and by K.-H. Yang, Am. J. Phys. 73, 742 (2005).9 The
9
http://kirkmcd.princeton.edu/examples/EM/jackson_ajp_70_917_02.pdf
http://kirkmcd.princeton.edu/examples/EM/yang_ajp_73_742_05.pdf
Princeton University 2001 Ph501 Set 8, Solution 2 23
potentials A(v) and V (v) in the so-called velocity gauge with the parameter v obey the
gauge condition,
c ∂V (v)
∇ · A(v) + 2 = 0. (95)
v ∂t
The scalar potential V (v) is obtained by replacing the speed of light c in the Lorenz-
gauge scalar potential by v. Equivalently, we replace the wave number k = ω/c by
k = ω/v. Thus, from eq. (85) we find,
(v) i(k R−ωt) ik 1
VE1 = −e 2
− 3 (P · R) (velocity gauge). (96)
R R
The vector potential in the v-gauge can be obtained from eq. (82) as,
(v) i i (v)
AE1 = − EE1 − ∇VE1
k k
i(kR−ωt)
e i(kR−ωt) 1 i
= −ik n̂ × (P × n̂) + e + [P − 3(P · n̂)n̂]
R R2 kR3
k 2 ei(k R−ωt) i(k R−ωt) k i
−i (P · n̂)n̂ − e + [P − 3(P · n̂)n̂]. (98)
k R kR2 kR3
This vector potential includes terms that propagate with velocity v both in the near
and far zones. When v = c, then k = k and the velocity-gauge vector potential (98)
reduces to the Lorenz-gauge potential (77); and when v → ∞, then k = 0 and the
velocity-gauge vector potential reduces to the Coulomb-gauge potential (91).10
cos(kR − ωt) sin(kR − ωt)
B = k P (n̂ × P̂)
2
− . (101)
R kR2
The Poynting vector contains six terms, some of which do not point along the radial
vector n̂:
c cos2 (kR − ωt) cos(kR − ωt) sin(kR − ωt)
S = k 4 P 2 [n̂ × (P̂ × n̂)] × (n̂ × P̂) −
4π R2 kR3
cos2(kR − ωt) − sin2 (kR − ωt)
+k 2 P 2 [3(P̂ · n̂)n̂ − P̂] × (n̂ × P̂)
R4
k 1
+ cos(kR − ωt) sin(kR − ωt) 3
−
R kR5
c cos2 (kR − ωt) cos(kR − ωt) sin(kR − ωt)
= 4 2
k P sin θn̂ 2
−
4π R2 kR3
cos2(kR − ωt) − sin2 (kR − ωt)
+k P [4 cos θP̂ + (3 cos θ − 1)n̂]
2 2 2
R4
k 1
+ cos(kR − ωt) sin(kR − ωt) 3
− , (102)
R kR5
where θ is the angle between vectors n̂ and P. As well as the expected radial flow
of energy, there is a flow in the direction of the dipole moment P. Since the product
cos(kR − ωt) sin(kR − ωt) can be both positive and negative, part of the energy flow
is inwards at times, rather than outwards as expected for pure radiation.
However, we obtain a simple result if we consider only the time-average Poynting
vector, S. Noting that cos (kR − ωt) = sin (kR − ωt) = 1/2 and
2 2
cos(kR − ωt) sin(kR − ωt) = (1/2) sin 2(kR − ωt) = 0, eq (102) leads to
ck 4 P 2 sin2 θ
S = n̂. (103)
8πR2
The time-average Poynting vector is purely radially outwards, and falls off as 1/R2
at all radii, as expected for a flow of energy that originates in the oscillating point
dipole (which must be driven by an external power source). The time-average angular
distribution d P /dΩ of the radiated power is related to the Poynting vector by
d P ck 4 P 2 sin2 θ P 2 ω 4 sin2 θ
= R2 n̂ · S = = , (104)
dΩ 8π 8πc3
which is the expression often quoted for dipole radiation in the far zone. Here we see
that this expression holds in the near zone as well.
We conclude that radiation, as measured by the time-average Poynting vector, exists
in the near zone as well as in the far zone.
equations in free space (i.e., where the charge density ρ and current density J are zero),
then the dual fields,
are solutions also. The Poynting vector is the same for the dual fields as for the original
fields,
c c
S = E × B = − B × E = S. (106)
4π 4π
Taking the dual of fields (10)-(11), we find the fields,
ei(kR−ωt) i
E = EM1 = −k 2 1+ n̂ × M, (107)
R kR
i(kR−ωt)
2e 1 i
B = BM1 = k n̂ × (M × n̂) + [3(n̂ · M)n̂ − M] 2 2 − . (108)
R k R kR
which are also solutions to Maxwell’s equations. These are the fields of an oscillating
point magnetic dipole, whose peak magnetic moment is M. In the near zone, the
magnetic field (108) looks like that of a (magnetic) dipole.
While the fields of eqs. (10)-(11) are not identical to those of eqs. (107)-(108), the
Poynting vectors are the same in the two cases. Hence, the time-average Poynting
vector, and also the angular distribution of the time-averaged radiated power are the
same in the two cases. The radiation of a point electric dipole is the same as that
of a point magnetic dipole (assuming that M = P), both in the near and in the far
zones. Measurements of only the intensity of the radiation could not distinguish the
two cases.
However, if measurements were made of both the electric and magnetic fields, then the
near zone fields of an oscillating electric dipole, eqs. (10)-(11), would be found to be
quite different from those of a magnetic dipole, eqs. (107)-(108). This is illustrated in
the figure on the previous page, which plots the ratio E/H = E/B of the magnitudes
of the electric and magnetic fields as a function of the distance r from the center of
the dipoles.
Princeton University 2001 Ph501 Set 8, Solution 2 26
To distinguish between the cases of electric and magnetic dipole radiation, it suffices
to measure only the polarization (i.e., the direction, but not the magnitude) of either
the electric of the magnetic field vectors.
Princeton University 2001 Ph501 Set 8, Solution 3 27
3. The rotating dipole p can be thought of as two oscillating linear dipoles oriented 90◦
apart in space, and phased 90◦ apart in time. This is conveniently summarized in
complex vector notation:
p = p0 (x̂ + iŷ) e−iωt , (109)
for a rotation from the +x̂ axis towards the +ŷ axis. Thus,
[p̈] × n̂ k 2 p0 e−i(kr−ωt)
Brad = = − (x̂ + iŷ) × n̂
c2 r r
k 2 p0 e−i(kr−ωt)
= (cos θ ŷ − i l̂), (111)
r
k 2 p0 e−i(kr−ωt)
Erad = Brad × n̂ = (cos θ l̂ + i ŷ). (112)
r
d P c 2 c 4 2
= r |Brad|2 = k p0 (1 + cos2 θ), (113)
dΩ 8π 8π
1 2 |p̈|2 2
P = 3
= 3 ω 4p20 , (115)
2 3c 3c
√ 2
since |p̈| = 2ω p0 in the present example.
Princeton University 2001 Ph501 Set 8, Solution 4 28
4. According to the Larmor formula, the rate of magnetic dipole radiation is,
dU 2 m̈2 2 m2 ω 4
= = , (116)
dt 3 c3 3 c3
where ω = 2π/T is the angular velocity, taken to be perpendicular to the magnetic
dipole moment m.
The radiated power (116) is derived from a decrease in the rotational kinetic energy,
U = Iω2 /2, of the pulsar:
dU 2
= −Iω ω̇ = MR2 ω |ω̇| , (117)
dt 5
where the moment of inertia I is taken to be that of a sphere of uniform mass density.
Combining eqs. (116) and (117), we have,
3 MR2 |ω̇| c3
m2 = . (118)
5 ω3
Substituting ω = 2π/T , and |ω̇| = 2π Ṫ /T 2 , we find,
3
2 3
m2 = MR T Ṫ c . (119)
20π 2
The static magnetic field B due to dipole m is,
3(m · r̂)r̂ − m
B= , (120)
r3
so the peak field at radius R is,
2m
B= . (121)
R3
Inserting this in eq. (119), the peak surface magnetic field is related by,
3
3 MT Ṫ c 3 (2.8 × 1033 )(7.5)(8 × 10−11 )(3 × 1010 )3
B2 = 2 = = 2.8 × 1030 gauss 2.
5π R4 5π 2 (106 )4
(122)
Thus, Bpeak = 1.7 × 1015 G = 38Bcrit , where Bcrit = 4.4 × 1013 G.
When electrons and photons of kinetic energies greater than 1 MeV exist in a magnetic
field with B > Bcrit, they rapidly lose this energy via electron-positron pair creation.
Kouveliotou et al. report that Bpeak = 8 × 1014 G without discussing details of their
calculation.
Princeton University 2001 Ph501 Set 8, Solution 5 29
5. The time-average field momentum density is given in terms of the Poynting vector as
(the real part of),
S c
Pfield = 2 = E × B . (123)
c 8π
Hence, the time-averaged angular momentum density is,
1 1
Lfield = r × Pfield = r × (E × B) = r[E(n̂ · B ) − B(n̂ · E)]. (124)
8πc 8πc
writing r = rn̂.
The time-average rate of radiation of angular momentum into solid angle dΩ is there-
fore,
d L 1 3
= cr2 Lfield = r [E(n̂ · B ) − B(n̂ · E)], (125)
dt dΩ 8π
since the angular momentum density L is moving with velocity c.
The radiation fields of an oscillating electric dipole moment p including both the 1/r
and 1/r2 terms of eqs. (79) and (87) are,
ei(kr−ωt) i 3i
E = k k+ p− k+ (n̂ · p)n̂ , (126)
r r r
ei(kr−ωt) i
B = k2 1+ (n̂ × p). (127)
r kr
Since n̂ · B = 0 for this case, only the second term in eq. (125) contributes to the
radiated angular momentum. We therefore find,
d L k3r i 2i ik 3
=− 1− (n̂ × p ) − (n̂ · p) = (n̂ · p)(n̂ × p ), (128)
dt dΩ 8π kr r 4π
ignoring terms in the final expression that have positive powers of r in the denominator,
as these grow small at large distances.
For the example of a rotating dipole moment (Prob. 2),
we have,
As we integrate over all directions of n̂, the contributions to d L /dt in the x-y plane
sum to zero, and only its z component survives. Hence,
1
d L d Lz k3 2k 3 2
= ẑ dΩ = 2π p20 ẑ sin2 θ d cos θ = p ẑ
dt dt dΩ 4π −1 3 0
2ck 3 2 P
= p ẑ = ẑ, (132)
3ω 0 ω
recalling eq. (115) for the radiated power P . Of course, the motion described by
eq. (129) has its angular momentum along the +z axis.
Princeton University 2001 Ph501 Set 8, Solution 6 31
7. Since the charge is assumed to rotate with constant angular velocity, the magnetic
moment it generates is constant in time, and there is no magnetic-dipole radiation.
Hence, we consider only electric-quadrupole radiation in addition to the electric-dipole
radiation. The radiated fields are therefore,
...
[p̈] × n̂ [Q] × n̂
B= + , E = B × n̂. (144)
c2 r 6c3 r
The electric dipole radiation fields are given by eqs. (111) and 112) when we write
p0 = ae.
The charge distribution is not azimuthally symmetric about any fixed axis, so we must
evaluate the full quadrupole tensor,
to find the components of the quadrupole vector Q. The position vector of the charge
has components,
ri = (a cos ωt, a sin ωt, 0), (146)
so the nonzero components of Qij are,
a2 e
Qxx = e(3x2 − r2 ) = a2e(3 cos2 ωt − 1) = (1 + 3 cos 2ωt), (147)
2
a2 e
Qyy = e(3y 2 − r2 ) = a2e(3 sin2 ωt − 1) = (1 − 3 cos 2ωt), (148)
2
Qzz = −er2 = −a2e, (149)
3a2e
Qxy = Qyx = 3exy = 3a2e sin ωt cos ωt = sin 2ωt. (150)
2
Only the time-dependent part of Qij contributes to the radiation, so we write,
⎛ ⎞
⎜ cos 2ωt sin 2ωt 0 ⎟
3a2 e ⎜
⎜
⎟
⎟
Qij (time dependent) = ⎜ sin 2ωt − cos 2ωt 0 ⎟ . (151)
2 ⎝ ⎜ ⎟
⎠
0 0 0
The unit vector n̂ towards the observer has components given in eq. (134), so the
time-dependent part of the quadrupole vector Q has components,
3a2 e
Qi = Qij nj = (cos 2ωt sin θ, sin 2ωt sin θ, 0). (152)
2
Thus,
... ...
[Qi ] = Qi (t = t − r/c) = −12a2 eω 3 sin θ(sin(2kr − 2ωt), cos(2kr − 2ωt), 0). (153)
It is preferable to express this vector in terms of the orthonormal triad n̂, ŷ, and
l̂ = ŷ × n̂, by noting that,
x̂ = n̂ sin θ − l̂ cos θ. (154)
Princeton University 2001 Ph501 Set 8, Solution 7 33
Hence,
...
[Q] = −12a2eω 3 sin θ(n̂ sin θ sin(2kr − 2ωt) − l̂ cos θ sin(2kr − 2ωt) + ŷ cos(2kr − 2ωt)).
(155)
The fields due to electric-quadrupole radiation are therefore,
...
[Q] × n̂ 2a2ek 3
BE2 = =− sin θ(l̂ cos(2kr − 2ωt) + ŷ cos θ sin(2kr − 2ωt)),(156)
6c3 r r
2a2ek 3
EE2 = BE2 × n̂ = sin θ(ŷ cos(2kr − 2ωt) − l̂ cos θ sin(2kr − 2ωt)). (157)
r
The angular distribution of the radiated power can be calculated from the combined
electric-dipole and electric-quadrupole fields, and will include a term ∝ k 4 due only
to dipole radiation as found in Prob. 2 above, a term ∝ k 6 due only to quadrupole
radiation, and a complicated cross term ∝ k 5 due to both dipole and quadrupole fields.
Here, we only display the term due to the quadrupole fields by themselves:
11
It is stated in eq. (71.5) of http://kirkmcd.princeton.edu/examples/EM/landau_ctf_75.pdf that, noting Lan-
dau’s Dαβ of his eq. (41.3) is our Qij ,
... 2
Qij
PE2 = . (159)
180c5
From the time-dependent part of the quadrupole tensor, eq. (151), we have,
⎛ ⎞
sin 2ωt − cos 2ωt 0
... ⎜ ⎟ ... 2
⎜ ⎟
Qij = 12eω3 a2 ⎜ − cos 2ωt − sin 2ωt 0 ⎟ , Qij = Q̈2ij = 288e2 ω6 a4 , (160)
⎝ ⎠ i,j
0 0 0
2e2 ω6 a4 8ca4 e2 k 6
PE2 = = , (161)
5c5 5
as in eq. (162).
Princeton University 2001 Ph501 Set 8, Solution 8 34
8. a) The dominant energy loss is from electric-dipole radiation, which obeys eq. (25),
dU 2a2 e2ω 4
= − PE1 = − . (163)
dt 3c3
For an electron of charge −e and mass m in an orbit of radius a about a fixed nucleus
of charge +e, F = ma tells us that
e2 v2
2
= m = mω 2 a, (164)
a a
so that,
e2
ω2 = , (165)
ma3
and the total energy (kinetic plus potential) is,
e2 1 2 e2
U = − + mv = − . (166)
a 2 2a
Using eqs. (165) and (166) in (163), we have,
dU e2 2e6
= 2 ȧ = − 4 2 3 , (167)
dt 2a 3a m c
or
1 da3 4e4 4
a2ȧ = = − 2 3 = − r02 c, (168)
3 dt 3m c 3
2 2
where r0 = e /mc is the classical electron radius. Hence,
With r0 = 2.8 × 10−13 cm and a0 = 5.3 × 10−9 cm, tfall = 1.6 × 10−11 s.
This is of the order of magnitude of the lifetime of an excited hydrogen atom, but the
ground state appears to have infinite lifetime in Nature.
This classical puzzle is pursued further in Prob. 9 below.
b) The analog of the quadrupole factor ea2 in Prob. 7 above for masses m1 and m2
in circular orbits with distance a between them is m1r12 + m2 r22 , where r1 and r2 are
measured from the center of mass. That is,
m 1 r1 = m 2 r2 , and r1 + r2 = a, (171)
so that,
m2 m1
r1 = a, r2 = a, (172)
m1 + m2 m1 + m2
Princeton University 2001 Ph501 Set 8, Solution 8 35
For the Earth-Sun system, a0 = 1.5 × 1013 cm, m1 = 6 × 1027 gm, m2 = 2 × 1033 cm,
and G = 6.7 × 10−10 cm2/(g-s2 ), so that tfall ≈ 1.5 × 1036 s ≈ 5 × 1028 years!
12
http://kirkmcd.princeton.edu/examples/GR/peters_pr_131_435_63.pdf
Princeton University 2001 Ph501 Set 8, Solution 9 36
9. The solution given here follows the succinct treatment by Landau, Classical Theory of
Fields, §74, http://kirkmcd.princeton.edu/examples/EM/landau_ctf_71.pdf
For charges in steady motion at angular frequency ω in a ring of radius a, the current
density J is periodic with period 2π/ω, so the Fourier analysis (34) at the retarded
time t can be evaluated via the usual approximation that r ≈ R − r · n̂, where R is
the distance from the center of the ring to the observer, r points from the center of
the ring to the electron, and n̂ is the unit vector pointing from the center of the ring
to the observer. Then,
[J] = J(r, t = t − r/c) = Jm (r ) e−imω(t−R/c+r ·n̂/c
m
im(kR−ωt) −imωr ·n̂/c
= e Jm (r ) e , (183)
m
where k = ω/c.
We first consider a single electron, whose azimuth varies as φ = ωt + φ0 , and whose
velocity is, of course, v = aω. The current density of a point electron of charge e can
be written using Dirac delta functions in a cylindrical coordinate system (ρ, φ, z) (with
volume element ρdρ dφ dz) as,
r = ρ(cos φ x̂+sin φ ŷ), n̂ = sin θ x̂+cos θ ẑ, and φ̂ = − sin φ x̂+cos φ ŷ. (186)
Using eqs. (185) and (186) in (183) and noting that ωT = 2π, we find,
ev im(kR−ωt) im(φ−φ0 −ωρ sin θ cos φ/c)
[J] = e e δ(ρ − a) δ(z) φ̂. (187)
2πρ m
The integrals yield Bessel functions with the aid of the integral representation (40).
The ŷ part of eq. (189) can be found by taking the derivative of this relation with
respect to z:
im+1 2π imφ−iz cos φ
Jm (z) = − e cos φ dφ, (190)
2π 0
For the x̂ part of eq. (189) we play the trick,
2π
0 = ei(mφ−z cos φ) d(mφ − z cos φ)
0
2π 2π
= m eimφ−iz cos φ dφ + z eimφ−iz cos φ sin φ dφ, (191)
0 0
so that,
1 2π imφ−iz cos φ m 1 2π imφ−iz cos φ m
e sin φ dφ = − e dφ = − m Jm (z). (192)
2π 0 z 2π 0 i z
Using eqs. (190) and (192) with z = mv sin θ/c in (189) we have,
ev im(kR−φ0 ) 1 1
Am = e m
Jm (mv sin θ/c) x̂ − m+1 Jm (mv sin θ/c) ŷ . (193)
cR i v sin θ/c i
We skip the calculation of the electric and magnetic fields from the vector potential,
and proceed immediately to the angular distribution of the radiated power according
to eq. (39),
dPm cR2 2 ck 2m2 R2 2
= |imk n̂ × Am | = |n̂ × Am |
dΩ 2π 2π
ck 2 m2R2 % 2 &
= cos θ |Am,x|2 + |Am,y |2
2π
ce2k 2 m2 2 2 v2 2
= cot θJm (mv sin θ/c) + 2 Jm (mv sin θ/c) . (194)
2π c
The present interest in this result is for v/c 1, but in fact it holds for any value of
v/c. As such, it can be used for a detailed discussion of the radiation from a relativistic
electron that moves in a circle, which emits so-called synchrotron radiation. This topic
is discussed further in Lecture 20 of the Notes.
We now turn to the case of N electrons uniformly spaced around the ring. The initial
azimuth of the nth electron can be written as,
2πn
φn = . (195)
N
The mth Fourier component of the total vector potential is simply the sum of compo-
nents (193) inserting φn in place of φ0:
N
ev im(kR−φn ) 1 1
Am = e m
Jm (mv sin θ/c) x̂ − m+1 Jm (mv sin θ/c) ŷ (196)
n=1 cR i v sin θ/c i
N
ev eimkR 1 1
= J m (mv sin θ/c) x̂ − J (mv sin θ/c) ŷ e−i2πmn/N .
cR im v sin θ/c im+1 m n=1
Princeton University 2001 Ph501 Set 8, Solution 9 38
This sum vanishes unless m is a multiple of N, in which case the sum is just N. The
lowest nonvanishing Fourier component has order N, and the radiation is at frequency
Nω. We recognize this as N th -order multipole radiation, whose radiated power follows
from eq. (194) as,
dPN ce2k 2 N 2 v2
= cot2 θJN2 (Nv sin θ/c) + 2 JN2 (Nv sin θ/c) . (197)
dΩ 2π c
For large N, but v/c 1, we can use the asymptotic expansion (41), and its derivative,
(ex/2)m
Jm (mx) ≈ √ (m 1, x 1), (198)
2πm x
to write eq. (197) as,
2N
dPN ce2 k 2N ev dPE1
≈ 2 2 sin θ (1 + cos2 θ) N (N 1, v/c 1). (199)
dΩ 4π sin θ 2c dΩ
In eqs. (198) and (199) the symbol e inside the parentheses is not the charge but rather
the base of natural logarithms, 2.718...
For currents in, say, a loop of copper wire, v ≈ 1 cm/s, so v/c ≈ 10−10 , while N ≈ 1023 .
The radiated power predicted by eq. (199) is extraordinarily small!
Note, however, that this nearly complete destructive interference depends on the elec-
trons being uniformly distributed around the ring. Suppose instead that they were
distributed with random azimuths φn . Then the square of the magnetic field at order
m has the form ,
2
N
−imφn
e−im(φl −φn ) = N.
2
|Bm | ∝
e
=N+ (200)
n=1 l =n
Thus, for random azimuths the power radiated by N electrons (at any order) is just
N times that radiated by one electron.
If the charge carriers in a wire were localized to distances much smaller than their
separation, radiation of “steady” currents could occur. However, in the quantum view
of metallic conduction, such localization does not occur.
The random-phase approximation is relevant for electrons in a so-called storage ring,
for which the radiated power is a major loss of energy – or source of desirable photon
beams of synchrotron radiation, depending on one’s point of view. We cannot expound
here on the interesting topic of the “formation length” for radiation by relativistic
electrons, which length sets the scale for interference of multiple electrons. See, for
example, http://kirkmcd.princeton.edu/accel/weizsacker.pdf
Princeton University 2001 Ph501 Set 8, Solution 10 39
10. We repeat the derivation of Prob. 1 above, this time emphasizing the advanced fields.
The advanced vector potential for the point electric dipole p = p0 e−iωt located at the
origin is,
We now consider the superposition of the fields (202)-(205) inside a conducting sphere
of radius a. The spatial part of the total electric field is then,
k2 3 ik
EE1 = − 3 (eikr − e−ikr ) + 3 2 (eikr + e−ikr ) (p0 · n̂) n̂
r r r
2
k 1 ikr −ikr ik ikr −ikr
− − 3 (e − e ) + 2 (e + e ) p0
r r r
k2 3 k
= 2i − 3 sin kr + 3 2 cos kr (p0 · n̂) n̂
r r r
2
k 1 k
−2i − 3 sin kr + 2 cos kr p0 . (206)
r r r
Princeton University 2001 Ph501 Set 8, Solution 10 40
Remarkably, this electric field is finite at the origin, although each of the fields (203)
and (205) diverges there. We also recognize that this electric field could be expressed
in terms of the so-called spherical Bessel functions,
sin x sin x cos x 3 1 3 cos x
j0 (x) = , j1 (x) = 2 − , j2 (x) = 3
− sin x− , ... (207)
x x x x x x2
An expansion of the spherical cavity field in terms of spherical Bessel functions occurs
“naturally” when we use the more standard approach to this problem, seeking solutions
to the Helmholtz wave equation via separation of variables in spherical coordinates.
See Electromagnetic Theory by J.A. Stratton (McGraw-Hill, 1941)13 for details of this
method.
Because the sum of the magnetic fields (202) and (204) is purely transverse, this cavity
mode is called a TM mode.
The boundary conditions at the surface of the sphere are that the radial component
of the magnetic field and the transverse component of the electric field must vanish.
Since the magnetic fields (202) and (204) are transverse at any radius, we examine the
electric field at r = a. Of the terms in eq. (206), only those in p0 have transverse
components, so the boundary condition is,
k k2 1
0 = 2 cos ka + sin ka − 3 , (208)
a a a
or,
1
cot ka = − ka, ⇒ ka = 2.744. (209)
ka
In case of a point magnetic dipole m = m0 e−iωt at the origin, the fields have the same
form as for an electric dipole, but with E and B interchanged. That is, the advanced
fields would be,
e−i(kr+ωt) i
EM 1,adv = k 2
−1 + n̂ × m0 , (210)
r
kr
−i(kr+ωt) k2 ik 1 k 2 ik 1
BM 1,adv = e − + 3 2 + 3 (m0 · n̂) n̂ + − 2 − 3 m0 . (211)
r r r r r r
and the retarded field due to magnetic dipole −m would be
i(kr−ωt)
2e i
EM 1,ret = −k 1+ n̂ × m0, (212)
r
kr
i(kr−ωt) k2 ik 1 k 2 ik 1
BM 1,ret = e + 3 2 − 3 (m0 · n̂) n̂ − + 2 − 3 m0 . (213)
r r r r r r
If the advanced and retarded magnetic dipole fields are superposed inside a spheri-
cal cavity of radius a, the condition that the transverse electric field vanish at the
conducting surface is,
ika i i sin ka
0=e 1+ + e−ika 1 − = 2 cos ka − , (214)
ka ka ka
13
http://kirkmcd.princeton.edu/examples/EM/stratton_electromagnetic_theory.pdf
Princeton University 2001 Ph501 Set 8, Solution 10 41
or,
tan ka = ka, ⇒ ka = 4.493. (215)
The electric field of this mode is purely transverse, so it is called a TE mode.
Clearly, other modes of a spherical cavity can be found by superposing the advanced
and retarded fields due to higher multipoles at the origin.
Princeton University 2001 Ph501 Set 8, Solution 11 42
11. This problem is due to D. Iwanenko and I. Pomeranchuk, On the Maximal Energy
Attainable in a Betatron, Phys. Rev. 65, 343 (1944).14
The electron is held in its circular orbit by the Lorentz force due to the magnetic field
B. Newton’s law, F = ma, for this circular motion can be written as,
γmv 2 v
F = γma = = e B. (216)
R c
For a relativistic electron, v ≈ c, so we have,
eRB
γ≈ . (217)
mc2
The electron is being accelerated by the electric field that is induced by the changing
magnetic flux. Applying the integral form of Faraday’s law to the circle of radius R,
we have (ignoring the sign),
Φ̇ πR2 Ḃave
2πREφ = = , (218)
c c
and hence,
RḂave
Eφ = , (219)
2c
The rate of change of the electron’s energy E due to Eφ is,
dE eRḂave
= F · v ≈ ecEφ = , (220)
dt 2
Since E = γmc2, we can write,
eRḂave
γ̇mc2 = , (221)
2
which integrates to,
eRBave
γ= . (222)
2mc2
Comparing with eq. (217), we find the required condition on the magnetic field:
Bave
B= . (223)
2
As the electron accelerates it radiates energy at rate given by the Larmor formula in
the rest frame of the electron,
dE 2e2p̈2 2e2a2
= − = − . (224)
dt 3c3 3c3
14
http://kirkmcd.princeton.edu/examples/EM/iwanenko_pr_65_343_44.pdf
Princeton University 2001 Ph501 Set 8, Solution 11 43
Because E and t are both the time components of 4-vectors, their transforms from
the rest frame to the lab frame have the same form, and the rate dE/dt is invariant.
However, acceleration at right angles to velocity transforms according to a = γ 2 a.
Hence, the rate of radiation in the lab frame is,
dE 2e2γ 4 a2 2e4γ 2 B 2
=− = − , (225)
dt 3c3 3m2 c3
using eq. (216) for the acceleration a.
The maximal energy of the electrons in the betatron obtains when the energy loss (225)
cancels the energy gain (220), i.e., when,
12. Since the dipole is much less than a wavelength away from the conducting plane, the
fields between the dipole and the plane are essentially the instantaneous static fields.
Thus, charges arranges themselves on the plane as if there were an image dipole at
distance d on the other side of the plane. The radiation from the moving charges on
the plan is effectively that due to the oscillating image dipole. A distant observer sees
the sum of the radiation fields from the dipole and its image.
The image dipole is inverted with respect to the original, i.e., the two dipoles are 180◦
out of phase.
Furthermore, there is a difference s in path length between the two dipole and the
distant observer at angles (θ, φ). We first calculate in a spherical coordinate system
with z axis along the first dipole, and x axis pointing from the plane to that dipole.
Then,the path difference is,
This path difference results in an additional phase difference δ between the fields from
the two dipoles at the observer, in the amount,
s 4πd
δ = 2π = sin θ cos φ. (229)
λ λ
If we label the electric fields due to the original and image dipoles as E1 and E2,
respectively, then the total field is,
dP |E|2 dP1
iδ 2
= = 1 − e A sin2 θ = 2A sin2 θ(1 − cos δ) = 4A sin2 θ sin2 δ/2
dΩ |E1 |2 dΩ
= 4A sin2 θ sin2 Δ, (231)
where,
δ 2πd
Δ= = sin θ cos φ. (232)
2 λ
Suppose we had chosen to use a spherical coordinate system (r, θ , φ) with the z axis
pointing from the plane to dipole 1, and the x axis parallel to dipole 1. Then, the
Princeton University 2001 Ph501 Set 8, Solution 12 45
sin2 θ = n2x + n2y = n2z + n2y = cos2 θ + sin2 θ cos2 φ = 1 − sin2 θ sin2 φ. (234)
If d = λ/4, then,
dP π
= 4A sin2 θ sin2 sin θ cos φ . (235)
dΩ 2
In the “side” view, φ = 0, so the pattern has shape,
π
sin2 θ sin2 sin θ (side view), (236)
2
while in the “top” view, θ = π/2 and the shape is,
2 π
sin cos φ (top view). (237)
2
This pattern has a single lobe in the forward hemisphere, as illustrated below:
This pattern does not radiate along the line from the plane to the dipole, as illustrated
below:
b) If the electric dipole is aligned with the line from the plane to the dipole, its image
has the same orientation.
The only phase difference between the radiation fields of the dipole and its image is
that due to the path difference δ, whose value has been given in eqs. (232) and (233).
It is simpler to use the angles (θ , φ ) in this case, since the radiation pattern of a single
dipole varies as sin2 θ. Then,
E = E1 + E2 = E1 (1 + eiδ ), (241)
dP |E|2 dP1 2
= 2 = 1 + eiδ A sin2 θ = 2A sin2 θ(1 + cos δ) = 4A sin2 θ cos2 δ/2
dΩ |E1| dΩ
= 4A sin2 θ cos2 Δ, (242)
with,
δ 2πd
Δ= = cos θ . (243)
2 λ
This radiation pattern is axially symmetric about the line from the plane to the dipole.
Princeton University 2001 Ph501 Set 8, Solution 12 47
If d = λ/4, then,
dP π
= 4A sin2 θ cos2 cos θ . (244)
dΩ 2
This pattern is a flattened version of the “donut” pattern sin2 θ , as illustrated below:
dP
= 4A sin2 θ cos2 (π cos θ) . (245)
dΩ
This pattern has a forward lobe for θ < π/6 and a “donut” for π/6 < θ < π/2, as
illustrated below:
c) For a magnetic dipole with axis parallel to the conducting plane, the image dipole
has the same orientation, the image consists of the opposite charge rotating in the
opposite direction, as shown below:
E = E1 + E2 = E1 (1 + eiδ ), (246)
Princeton University 2001 Ph501 Set 8, Solution 12 48
dP |E|2 dP1
iδ 2 2 2 2 2
= = 1 + e A sin θ = 2A sin θ(1 + cos δ) = 4A cos θ sin δ/2
dΩ |E1 | dΩ
2
where,
δ 2πd
Δ= = sin θ cos φ. (248)
2 λ
If d = λ/4, then,
dP π
= 4A sin2 θ cos2 sin θ cos φ . (249)
dΩ 2
In the “side” view, φ = 0, so the pattern has shape,
2 2 π
sin θ cos sin θ (side view), (250)
2
while in the “top” view, θ = π/2 and the shape is,
π
cos2 cos φ (top view). (251)
2
This pattern, shown below, is somewhat similar to that of part a) for d = λ/2.
dP
= 4A sin2 θ cos2 (π sin θ cos φ) . (252)
dΩ
In the “side” view, φ = 0, so the pattern has shape,
This pattern, shown below, is somewhat similar to that of part b) for d = λ/2.
Princeton University 2001 Ph501 Set 8, Solution 12 49
Finally, we consider the case of a magnetic dipole aligned with the line from the plane
to the dipole, in which case its image has the opposite orientation.
As in part b), the only phase difference between the radiation fields of the dipole and
its image is that due to the path difference δ, whose value has been given in eqs. (232)
and (233). We use the angles (θ, φ ) in this case, since the radiation pattern of a single
dipole varies as sin2 θ. Then,
dP |E|2 dP1
iδ 2 2 2 2
= = 1 − e A sin θ = 2A sin θ (1 − cos δ) = 4A sin θ sin δ/2
2
dΩ |E1|2 dΩ
= 4A sin2 θ sin2 Δ, (256)
with,
δ 2πd
Δ= = cos θ . (257)
2 λ
This radiation pattern is axially symmetric about the line from the plane to the dipole.
If d = λ/4, then,
dP π
= 4A sin2 θ sin2 cos θ . (258)
dΩ 2
This pattern, shown below, is somewhat similar to that of part a) for d = λ/2.
Princeton University 2001 Ph501 Set 8, Solution 12 50
dP
= 4A sin2 θ sin2 (π cos θ ) . (259)
dΩ
This pattern is qualitatively similar to that for d = λ/4, shown just above, but the
maximum occurs at a larger value of θ .
Princeton University 2001 Ph501 Set 8, Solution 13 51
13. From p. 191, Lecture 16 of the Notes we recall that a single, short, center-fed, linear
antenna of dipole moment,
I0L e−iωt
p(t) = i (260)
2ω
radiates time-averaged power (according to the Larmor formula),
For the record, the current distribution in this short antenna is well approximated by
the triangular waveform,
−iωt 2 |z|
I(z, t) = I0 e 1− . (262)
L
The associated charge distribution ρ(z, t) is related by charge conservation, ∇·J = −ρ̇,
which for a 1-d distribution is simply,
∂I 2
ρ̇ = − = −I0 e−iωt ∓ , (263)
∂z L
so that,
2iI0 e−iωt
ρ=± , (264)
ωL
and the dipole moment is given by,
L/2
I0L e−iωt
p= ρz dz = i , (265)
−L/2 2ω
as claimed above.
Turning to the case to two antennas, we proceed as in the previous problem and write
their combined electric field as,
where now the phase difference δ has contributions due to the path difference for
radiation from the two antennas (whose separation is d = λ/4), as well as from their
intrinsic phase difference of 90◦ . That is,
2π λ π π
δ= cos θ + = (1 + cos θ). (267)
λ 4 2 2
From eqs. (261), (266) and (267) we find
dU dU1
iδ 2 ω 2 I02L2 sin2 θ
= 1 − e = (1 − cos δ)
dtdΩ dtdΩ 16πc3
ω 2 I02L2 sin2 θ π
= 1 + sin cos θ . (268)
16πc3 2
Princeton University 2001 Ph501 Set 8, Solution 13 52
This angular distribution favors the forward hemisphere, as shown in the sketch:
14. According to p. 181, Lecture 16 of the Notes, the power radiated from a known current
distribution that oscillates at angular frequency ω is given by,
2
dPω 1 −ik·r
= Jω (r ) × k e dVol , (272)
dΩ 8πc
where k = n̂ ω/c.
We take the z-axis along the antenna, so the radiated power due to the current distri-
bution,
2πz −iωt
I(z) = I0 sin e (−L/2 < z < L/2), (273)
L
is,
2
dPω 1 L/2 2πz
= I0 sin k sin θ e−ikz cos θ dz
dΩ 8πc −L/2 L
2
2 2
k I0
= sin2 θ , (274)
8πc
where,
L/2
2πz
= sin [cos(kz cos θ) − i sin(kz cos θ)] dz. (275)
−L/2 L
For L = λ, we have k = 2π/L and,
π/k
= sin kz[cos(kz cos θ) − i sin(kz cos θ)] dz. (276)
−π/k
we have,
1 π
Im = {[cos[x(1 + cos θ)] − cos[x(1 − cos θ)]} dx.
k 0
1 sin[π(1 + cos θ)] sin[π(1 − cos θ)]
= −
k 1 + cos θ 1 − cos θ
1 sin(π cos θ) sin(π cos θ)
= − −
k 1 + cos θ 1 − cos θ
2 sin(π cos θ)
= . (279)
k sin2 θ
Princeton University 2001 Ph501 Set 8, Solution 14 54
To cast this in the form of a known special function, we let 1 + u = v/2π, so that,
4π
I02 1 − cos v I2 I2
P = dv = 0 Cin(4π) = 3.11 0 , (282)
2c 0 v 2c 2c
where Cin is the so-called cosine integral.
b) To calculate the radiation in the multipole approximation, we need to convert the
current distribution I(z) e−iωt to a charge distribution ρ(z, t). This is accomplished via
the continuity equation,
∂I
= −ρ̇ = iωρ. (283)
∂z
For the current distribution (273) we find,
2πi 2πz −iωt
ρ=− I0 cos e . (284)
ωL L
The dipole moment of this distribution is,
L/2
p= ρz dz = 0, (285)
−L/2
so there is no electric-dipole radiation. As the current flows along a line, and not in a
loop, there is no magnetic-dipole radiation either.
Princeton University 2001 Ph501 Set 8, Solution 14 55
The charge distribution is symmetric about the z axis, so its tensor quadrupole moment
can be characterized in terms of the single quantity,
L/2 L/2 L/2
8πi 2πz
Qzz = 2
2ρz dz = 4 ρz dz = −
2
I0 z 2 cos dz
−L/2 0 ωL 0 L
π
L2 i 2L2 i
= − 2 I0 2
x cos x dx = I0 , (286)
π ω 0 πω
using the integral (55).
The total power radiated by the symmetric quadrupole moment is, according to p. 190,
Lecture 16 of the Notes,
|Qzz |2 ω 6 L4 ω 4 I02
PE2 = = . (287)
240c5 30π 2 c4 2c
When L = λ = 2πc/ω, this becomes,
8π 2 I02 I2
PE2 = = 5.26 0 . (288)
15 2c 2c
In this example, higher multipoles must contribute significantly to the total power,
reducing it to the “exact” result (282).
Princeton University 2001 Ph501 Set 8, Solution 15 56
where in the dipole approximation, the far-zone scattered electric field is,
ei(kr−ωt)
Escat = k 2 [(n̂ × p0 ) × n̂ − n̂ × m0 ] , (290)
r
and p0 eiωt and m0 e−iωt are the electric- and magnetic-dipole moments induced in the
conducting sphere by the incident wave.
Because the incident wavelength is large compared to the radius of the sphere, the
incident fields are essentially uniform over the sphere, and the induced fields near the
sphere are the same as the static fields of a conducting sphere in an otherwise uniform
electric and magnetic field. Then, from p. 57, Lecture 5 of the Notes, the induced
electric-dipole moment is given by
p0 = a3 E0. (291)
For the induced magnetic dipole, we recall p. 98, Lecture 8 of the Notes, remembering
that a conducting sphere can be thought of a permeable sphere with zero permeabil-
ity and a dielectric sphere of infinite dielectric constant. Hence, the magnetic-dipole
moment is,
a3
m 0 = − B0 . (292)
2
Then,
ei(kr−ωt) 1
Escat = −k 2a3 n̂ × (E0 × n̂) + (n̂ × B0 , (293)
r 2
where n̂ is along the vector r that points from the center of the sphere to the distant
observer.
For a wave propagating in the +z direction with electric field linearly polarized along
direction l̂, E0 = E l̂, and the magnetic field obeys B0 = ẑ × E0 . Then,
ei(kr−ωt) 1
Escat = −k a E0
2 3
n̂ × (l̂ × n̂) + n̂ × (ẑ × l̂)
r
2
i(kr−ωt)
e (n̂ · ẑ) ẑ
= −k 2 a3E0 l̂ 1 − + n̂ − (n̂ · l̂) . (294)
r 2 2
For an observer in the x-z plane, n̂· ẑ = cos θ. Then, for electric polarization parallel to
the scattering plane n̂ · l̂ = sin θ, while for polarization perpendicular to the scattering
plane n̂ · l̂ = 0.
From eqs. (296) we see that dσ ⊥ /dΩ is always nonzero, but dσ /dΩ = 0 for θ = π/3,
so for this angle, the scattered radiation is linearly polarized parallel to the scattering
plane for arbitrary incident polarization.
Addendum: The Fields and Poynting Vector Close to the Sphere
Using the results of Prob. 2 above we can also discuss the fields close to the sphere.
In particular, from eqs. (11) and (107) the scattered electric field at any position r
outside the sphere is,
ei(kr−ωt) 1 i
Escat(r, t) = k 2 (n̂ × p0) × n̂ + [3(n̂ · p0 ) n̂ − p0] 2 2 −
r k r kr
Princeton University 2001 Ph501 Set 8, Solution 15 58
i
− 1+ n̂ × m0
kr
i(kr−ωt)
2 3e 1 i
= k a (n̂ × E0) × n̂ + [3(n̂ · E0 ) n̂ − E0 ] 2 2 −
r k r kr
1 i
+ 1+ n̂ × B0 , (299)
2 kr
also using eqs. (291) and (292). In this Addendum, we suppose that the electric field of
the incident plane wave is along the x-axis, so that E0 = E x̂ and B0 = E ŷ, while the
point of observation is at r = (r, θ, φ). We express the electric-field vector in spherical
coordinates, noting that,
n̂ = r̂, (300)
x̂ = sin θ cos φ r̂ + cos θ cos φ θ̂ − sin φ φ̂, (301)
ŷ = sin θ sin φ r̂ + cos θ sin φ θ̂ + cos φ φ̂, (302)
ẑ = cos θ r̂ − sin θ θ̂. (303)
Thus,
ei(kr−ωt) '
Escat(r, t) = k 2 a3E0 cos θ cos φ θ̂ − sin φ φ̂
r
1 i
+(2 sin θ cos φ r̂ − cos θ cos φ θ̂ + sin φ φ̂) 2 2 −
k r kr
1 i
− 1+ (cos φ θ̂ − cos θ sin φ φ̂)
2 kr
ei(kr−ωt) 1 i
2 3
= k a E0 2 sin θ cos φ − r̂
r k 2r2 kr
1 i 1 i
+ cos φ cos θ 1 − 2 2 + − 1+ θ̂
k r kr 2 kr
1 i cos θ i
− sin φ 1 − 2 2 + − 1+ φ̂ . (304)
k r kr 2 kr
Similarly, using eqs. (8) and (108) the scattered magnetic field can be written as,
ei(kr−ωt) 1 i
Bscat(r, t) = k 2
(n̂ × m0 ) × n̂ + [3(n̂ · m0) n̂ − m0 ] 2 2 −
r k r kr
i
+ 1+ n̂ × p0
kr
i(kr−ωt)
2 3 e 1 i
= −k a (n̂ × B0 ) × n̂ + [3(n̂ · B0 ) n̂ − B0] 2 2 −
2r k r kr
i
−2 1 + n̂ × E0
kr
ei(kr−ωt) '
= −k 2 a3E0 cos θ sin φ θ̂ + cos φ φ̂
2r
1 i
, +(2 sin θ sin φ r̂ − cos θ sin φ θ̂ − cos φ φ̂) 2 2 −
k r kr
Princeton University 2001 Ph501 Set 8, Solution 15 59
i
−2 1 + (sin φ θ̂ + cos θ cos φ φ̂)
kr
ei(kr−ωt) 1 i
= −k a E0
2 3
2 sin θ sin φ − r̂
2r k 2r2 kr
1 i i
+ sin φ cos θ 1 − 2 2 + −2 1+ θ̂
k r kr kr
1 i i
+ cos φ 1 − 2 2 + − 2 cos θ 1 + φ̂ . (305)
k r kr kr
On the surface of the sphere, r = a, the scattered electromagnetic fields are, to the
leading approximation when ka 1,
on the surface of the sphere is purely radial, and the total magnetic field,
3
Btot(r = a) = Bin(r = a) + Bscat(r = a) = E0 e−iωt (cos θ sin φ θ̂ + cos φ φ̂), (311)
2
is purely tangential, as expected for a perfect conductor.
The total charge density σtot on the surface of the conducting sphere follows from
Gauss’ law as,
Etot(r = a) · r̂ 3E0 −iωt 3
σ tot = = e sin θ cos φ = σ scat, (312)
4π 4π 2
where σ scat is the surface charge density corresponding to the scattered field (306).
Similarly, the total current density Ktot on the surface of the sphere follows from
Ampère’s law as,
c 3cE0 −iωt
Ktot = r̂ × Btot(r = a) = e (− cos φ θ̂ + cos θ sin φ φ̂) = 3 Kscat , (313)
4π 8π
where Kscat is the surface charge density corresponding to the scattered field (307).
We can now discuss the energy flow in the vicinity of the conductor sphere from two
perspectives. These two views have the same implications for energy flow in the far
zone, but differ in their description of the near zone.
Princeton University 2001 Ph501 Set 8, Solution 15 60
First, we can consider the Poynting vector constructed from the total electromagnetic
fields,
c
Stot = Etot × Btot. (314)
4π
Because the tangential component of the total electric field vanishes at the surface of
the sphere, lines of the total Poynting vector do not begin or end on the sphere, but
rather they pass by it tangentially. In this view, the sphere does not absorb or emit
energy, but simply redirects (scatters) the flow of energy from the incident wave.
However, this view does not correspond closely to the “microscopic” interpretation that
atoms in the sphere are excited by the incident wave and emit radiation as a result,
thereby creating the scattered wave. We obtain a second view of the energy flow that
better matches the “microscopic” interpretation if we write,
c
Stot = Etot × Btot
4π
c
= (Ein + Escat) × (Bin + Bscat)
4π
c c c
= Ein × Bin + (Ein × Bscat + Escat × Bin) + Escat × Bscat
4π 4π 4π
= Sin + Sinteraction + Sscat. (315)
Since the scattered fields (306)-(307) at the surface of the sphere include tangential
components for both the electric and the magnetic field, the scattered Poynting vector,
Sscat, has a radial component, whose time average we wish to interpret as the flow of
energy radiated by the sphere. The scattered Poynting vector at any r is given by,
c
Sscat = Re(Escat × Bscat)
8π (
c
= Re (Eθ,scat Bφ,scat − Eφ,scat Bθ,scat) r̂ + (Eφ,scat Br,scat − Er,scat Bφ,scat) θ̂
8π )
+(Er,scat Bθ,scat − Eθ,scat Br,scat) φ̂
⎧⎡ 2 ⎤
2
c k 4 a6E02 ⎨⎣ 2 1 cos θ ⎦
= cos φ − cos θ + sin φ 1 −
2
r̂
8π r2 ⎩ 2 2
1 cos θ
− 4 4 r̂ + sin θ θ̂ . (316)
k r 2
The radial term of eq. (316) in square brackets is identical to the far-zone Poynting
vector. However, close to the sphere we find additional terms in Sscat, so that in the
near zone Srad = Sscat. Indeed, at the surface of the sphere we find
⎧ ⎡ 2 ⎤
2
c 2⎨ 4 4⎣ 2 1 cos θ ⎦
Sscat(r = a) = E0 k a cos φ − cos θ + sin2 φ 1 − r̂
8π ⎩ 2 2
cos θ
− r̂ + sin θ θ̂
2
c 2 cos θ
≈ − E0 r̂ + sin θ θ̂ . (317)
8π 2
Princeton University 2001 Ph501 Set 8, Solution 15 61
Of course, the conducting sphere is not an energy source by itself, and the radiated
energy is equal to the energy absorbed from the incident wave. For a description of
the flow of energy that is absorbed, we look to the time-average of the incident and
interaction terms of eq. (315). Lines of the incident Poynting vector,
c , c 2
Sin = E0 ẑ = E (cos θ r̂ − sin θ θ̂), (318)
8π 8π 0
enter and leave the sphere with equal strength, and are therefore not to be associated
with energy transfer to the radiation fields. So, we look to the interaction term,
c (
Sinteraction = Re (Eθ,scat Bφ,in + Eθ,in Bφ,scat − Eφ,scat Bθ,in − Eφ,in Bθ,scat) r̂
8π
+(Eφ,scat Br,in + Eφ,in Br,scat − Er,scat Bφ,in − Er,in Bφ,scat), θ̂
)
+(Er,scat Bθ,in + Er,in Bθ,scat − Eθ,scat Br,in − Eθ,in Br,scat) φ̂
c k 2a3 E02 cos θ
= − cos[kr(1 − cos θ)]
8π r 2k 2r2
1 cos θ
+(1 + cos θ) cos φ 2
− cos θ + sin2 φ −1 ×
2 2
sin[kr(1 − cos θ)]
− cos[kr(1 − cos θ)] r̂
kr
sin θ 9
+ cos[kr(1 − cos θ)] 2 2 2 − cos φ + . . . θ̂
2
k r 2
9 sin 2θ sin 2φ
+ cos[kr(1 − cos θ)] + . . . φ̂ , (319)
8 k 2 r2
where the omitted terms are small close to the sphere. Note that in the far zone
the time-average interaction Poynting vector contains terms that vary as 1/r times
cos[kr(1 − cos θ)]. These large terms oscillate with radius r with period λ, and might
be said to describe a radial “sloshing” of energy in the far zone, rather than a radial
flow. It appears in practice that one cannot detect this “sloshing” by means of a
small antenna placed in the far zone, so we consider these terms to be unphysical.
Nonetheless, it is interesting that they appear in the formalism.
At the surface of the sphere we have, again for ka 1,
c 2 cos θ 9 9
Sinteraction(r = a) = E0 − r̂ + sin θ 2 − cos2 φ θ̂ + sin 2θ sin 2φ φ̂ .
8π 2 2 8
(320)
The total Poynting vector on the surface of the sphere is the sum of eqs. (317), (318)
and (320),
c 2 9 9
Stot (r = a) = E0 − sin θ cos2 φ θ̂ + sin 2θ sin 2φ φ̂ . (321)
8π 2 8
The radial component of the total Poynting vector vanishes on the surface of the sphere,
as expected for a perfect conductor.
Princeton University 2001 Ph501 Set 8, Solution 15 62