Radiation
Radiation
In the previous lectures, we found that particles interacting with scattering centres (i.e. other
particles or fields) can be accelerated to very high energies. When these accelerated particles
are charges, they produce electromagnetic waves. Radiation is an irreversible flow of
electromagnetic energy from the source (charges) to infinity. This is possible only because the
electromagnetic fields associated with accelerating charges fall off as 1/r instead of 1/r 2 as
is the case for charges at rest or moving uniformly. So the total energy flux obtained from the
Poynting flux is finite at infinity.
Note: The material in this lecture is largely a review of some of the content presented in the
Advanced Electromagentic Theory courses in 3rd and 4th years. As such, it is
non-examinable for this course.
40
! |r(t) − r0 (t! )|
t =t− (1)
c
41
PHYS 4011 – HEA Lec. 5
The radiation field at P at time t is calculated from the retarded scalar and vector potentials V
and A using E = −∇V − ∂A/∂t and B = ∇ × A. But Maxwell’s equations define these
in terms of the continuous charge and current densities ρ and J. So to evaluate V and A for a
point charge, it is necessary to integrate over the volume distribution at one instant in time
taking the limit as the size of the volume goes to zero. Formally, this can be done by taking
ρ(r, t) = qδ(r − r0 (t)) and J(r, t) = qv(t)δ(r − r0 (t)), where v(t) = ṙ0 (t) is the
velocity of the charge. Similarly, another delta function is introduced to single out only the
source point at the retarded time that we are interested in. After integrating over the volume,
we have
! ! ! ! R(t! )
R(t ) = r(t) − r0 (t ) , R(t ) = |r(t) − r0 (t )| , R̂ =
R(t! )
42
µ0 qv(t! ) v
A(r, t) = = 2 V (r, t) Liénard–Wiechart potentials (3)
4π (1 − R̂ · v(t! )/c)R c
These are the famous retarded potentials for a moving point charge.
Points to note:
1. The factor (1 − R̂ · v(t! )/c) implies geometrical beaming. It means that the potentials
are strongest at field points lying ahead of the source point S ! and closely aligned with the
particle’s trajectory. The effect is enhanced when the particle speed becomes relativistic.
2. Retardation is what makes it possible for a charged particle to radiate. To see why,
note that the potentials fall off as 1/r . Differentiation to retrieve the fields would yield a
1/r 2 fall-off if there were no other r -dependence in the potentials. This does not give rise
to a net electromagnetic energy flux as r → ∞ and hence, no radiation field. (Recall that
!
the rate of change of EM energy goes as S · dA). However, there is an implicit
r -dependence in the retarded time that leads to a 1/r -dependence in the fields, upon
differentiation of the potentials. This does result in a net flow of EM energy towards infinity.
43
PHYS 4011 – HEA Lec. 5
The differentiation of the Liénard–Weichart potentials to obtain the radiation field of a single
moving charge is lengthly, but straightforward (see Jackson for details). Writing the charged
particle’s velocity at the retarded time as βc = ṙ0 (t! ) and its corresponding acceleration as
β̇c = r̈0 (t! ), the fields are
# &
q (R̂ − β)(1 − β 2 ) R̂ 1$ %
E(r, t) = + × (R̂ − β) × β̇ (4)
4π%0 (1 − R̂ · β)3 R2 (1 − R̂ · β)3 R c
1
B(r, t) = R̂ × E(r, t) (5)
c
The first term on the RHS of the E-field is the velocity field. It falls off as 1/R2 and is just the
generalisation of Coulomb’s Law to uniformly moving charges. The second term is the
acceleration field contribution. It falls off as 1/R, is proportional to the particle’s acceleration
and is perpendicular to R̂.
44
This electric field, along with the corresponding magnetic field, constitute the radiation field of a
moving charge:
# &
q R̂ 1$ %
Erad (r, t) = × (R̂ − β) × β̇
4π%0 (1 − R̂ · β)3 R c
1
Brad (r, t) = R̂ × Erad (r, t) (6)
c
Note that Erad , Brad and R̂ are mutually perpendicular.
45
PHYS 4011 – HEA Lec. 5
and Brad (r, t) follows from (6). Note that Erad lies in the plane containing R̂ and v̇ (i.e. the
plane of polarisation) and Brad is perpendicular to this plane. If we let θ be the angle between
R̂ and v̇, then
q v̇
|Erad | = c|Brad | = sin θ
4π%0 Rc2
The Poynting vector is in the direction of R̂ and has the magnitude
1 2 µ0 q 2 v̇ 2
S= Erad = sin2 θ (8)
µ0 c 16π 2 c R2
We now want to express this as an emission coefficient.
46
Since S is the EM energy dW emitted per unit time dt per unit area dA (i.e.
S = dW/(dtdA)), we can write dA = R2 dΩ , where dΩ is the solid angle about the
direction R̂ of S. So the power emitted per solid angle is
dW µ0 2 2 2
= S R2 = q v̇ sin θ
dtdΩ 16π 2 c
Note the characteristic dipole pattern ∝ sin2 θ : there is no emission in the direction of
acceleration and the maximum radiation is emitted perpendicular to the acceleration. The total
electromagnetic power emitted into all angles is obtained by integrating this:
2π π +1
dW µ0 2 2 µ0 2 2
" " "
2
P = = q v̇ sin θ sin θ dθ dφ = q r̈0 (1 − µ2 ) dµ
dt 16π 2 c 0 0 8πc −1
The integral gives a factor of 4/3, whence we arrive at the Larmor formula for the
electromagnetic power emitted by an accelerating charge:
µ0 q 2 r̈02
P = Larmor formula (9)
6πc
47
PHYS 4011 – HEA Lec. 5
To calculate the radiation field from a system of many moving charges, we must keep track of
the phase relations between the radiating sources, because the retarded times will differ for
each charge. In some situations, however, it is possible to neglect this complication and use
the principle of superposition to determine the properties of the radiation field at large r .
48
where Ri is the distance between each source point ri and the field point r . But the
differences between the Ri are negligible, particularly as r → ∞. So we can just keep
R = |r − r0 (t! )| as a characteristic distance and use the definition for the dipole moment,
viz.
+
d= qi r i (10)
i
to get
' *
1 R̂ 1 ( )
Erad = × 2 R̂ × d̈ (11)
4π%0 R c
Then, as before, we find
dW µ0 ¨2 2
= d sin θ
dtdΩ 16π 2 c
49
PHYS 4011 – HEA Lec. 5
50
which is valid provided v̇(t) is real. Applying these realtions to the Larmor formula (9), we can
51
PHYS 4011 – HEA Lec. 5
determine the total energy radiated by a single charged particle with an acceleration history
v̇(t):
+∞ +∞
µ0 q 2 µ0 q 2 ∞
" " "
2
P dt = |v̇(t)| dt = |v̇(ω)|2 dω (16)
−∞ −∞ 6πc 3πc 0
!∞
Since the total emitted energy must also equal 0 (dW/dω) dω , then the energy per unit
bandwidth is just
dW µ0 q 2
= |v̇(ω)|2 (17)
dω 3πc
Clearly, this energy is just emitted during the period that the particle is experiencing
acceleration. In the dipole approximation for multiple particles, this expression becomes
dW µ0 4
= ω |d(ω)|2 (18)
dω 3πc
This is because differentiating wrt t twice introduces a factor ω 2 in the Fourier transform – c.f.
! +∞
d(t) = (2π)−1/2 −∞ d(ω) exp(−iωt) dω =⇒
¨ = −(2π)−1/2 +∞ ω 2 d(ω) exp(−iωt) dω , so ev̇(ω) = ω 2 d(ω).
!
d(t) −∞
52
PHYS 4011 – HEA Lec. 6
When a high speed electron encounters the Coulomb field of another charge, it emits
bremsstrahlung radiation, also known as free-free emission. The word bremsstrahlung means
braking radiation because the electron rapidly decelerates when the other charge is a massive
ion. The derivation can be done classically using the dipole approximation for nonrelativistic
particles, with quantum corrections added as “Gaunt factors” to the classical formulas. The
quantum corrections become important when photon energies become comparable to
energies of the emitting particles. We only need to consider electron-ion bremsstrahlung
because for collisions between like charges (e.g. electron-electron), the dipole approximation
predicts zero radiation and a higher order calculation is required. This also means that less
radiation is emitted for collisions between like particles. In electron-ion bremsstrahlung, the
electrons are the primary emitters because their accleration is ∼ mp /me times greater.
54
Let R be the position vector of the electron from the ion. Then the dipole moment is
d = −eR and its second time derivative is d̈ = −ev̇. We want an emission spectrum for the
bremsstrahlung radiation using eqn. (17) in Lec. 5, viz.
dW µ0 e2
= |v̇(ω)|2 (1)
dω 3πc
and using
+∞
1
!
v̇(ω) = v̇(t) exp(iωt) dt (2)
(2π)1/2 −∞
55
PHYS 4011 – HEA Lec. 6
Consider first that the electron is interacting with the ion only over a finite collision time
τ # b/v . When ωτ $ 1 the exponential in the integrand oscillates rapidly and the resulting
integral is small. When ωτ % 1, on the other hand, the exponential term is approximately
unity and the resulting integral is just v̇ = dv/dt ≈ ∆v over the time interval dt ≈ τ . So we
have
1
(2π)1/2
∆v ωτ % 1
|v̇(ω)| # (3)
0 ωτ $ 1
So our radiation spectrum goes as
µ e2
2
6π 2 c |∆v| b % v/ω
0
dW
# (4)
dω 0 b $ v/ω
56
Now we can work out ∆v by noting that the total Coloumb force on the electron is
Ze2 /(4π$0 R2 ) in the −R̂ direction. The perpendicular component of acceleration is the
strongest and will thus make the dominant contribution to the radiation spectrum, so we can
write
1 Ze2 b dt 1 Ze2 2
!
∆v # = (5)
4π$0 me (b2 + v 2 t2 )3/2 4π$0 me bv
(the integral turns out to be elementary). Substituting this into the expression for the radiation
spectrum gives
2 6
8Z e
dW
3πc3 (4π"0 )3 m2e b2 v 2 b % v/ω
# (6)
dω 0 b $ v/ω
This is the spectrum for small angle scatterings by a single electron off a single ion. Next we
want to generalise this to the case of a realistic plasma in which we have many electrons
interacting with many ions.
57
PHYS 4011 – HEA Lec. 6
Let the ion and electron number densities in the plasma be ni and ne . Then the flux of
electrons incident on an ion is ne v for a fixed electron speed v . The element of area about an
ion over which an electron encounter occurs is approximately 2πb db. So the emission per unit
time per unit volume per unit frequency range is
dW ∞
dW (b)
!
= ne ni 2πv b db (7)
dωdV dt bmin dω
where bmin is a minimum impact parameter to be chosen. Now it is difficult to see how the
solution for dW/dω obtained above in the asymptotic limits b % v/ω and b $ v/ω can be
used to solve this integral over a full range of impact parameters. However, it turns out that the
solution can be well approximated by using the non-zero asymptotic solution for dW/dω
because the integral is just logarithmic in b:
16ne ni Z 2 e6
% &
dW bmax
# 3 ln (8)
dωdV dt 3c (4π$0 )3 m2e v bmin
where bmax ∼ v/ω is some value beyond which the b % v/ω limit no longer applies and the
contribution to the integral becomes negligible. We can set bmax = v/ω , even though it is
58
uncertain because it is inside the logarithm. An appropriate value of bmin can be chosen to
correspond to the break down of the small-angle scattering approximation. So when ∆v ∼ v,
(5) implies bmin # 2Ze2 /(4π$0 me v 2 ).
The exact expression for the radiation spectrum can be obtained with a full quantum treatment.
For convenience, a quantum correction is added to the classical formula. This correction term
is known as the Gaunt factor,
√ % &
3 bmax
Gff (v, ω) = ln free-free Gaunt factor (9)
π bmin
giving
dW 16πne ni Z 2 e6
# 3/2 3 Gff (v, ω) (10)
dωdV dt 3 c (4π$0 )3 m2e v
This is now the bremsstrahlung radiation emitted per unit time per unit volume per unit
frequency by single-speed electrons interacting with many ions. Next we compute the volume
emissivity for a thermal distribution of electron speeds.
59
PHYS 4011 – HEA Lec. 6
mv 2
% &
3 2
dn(v) = f (v)d v = f (v)4πv dv ∝ exp − v 2 dv
2kT
We need to average the single-speed radiation spectrum over this distribution function for all
electron speeds satisfying 12 me v 2 >
∼ hω/2π , i.e.
dW (v,ω) 2
v exp(−mv 2 /2kT ) dv
'∞
dW (T, ω) vmin dωdV dt
= '∞ (11)
dωdV dt 0
v 2 exp(−mv 2 /2kT ) dv
Using dω = 2π dν , the final result is the expression for the free-free volume emissivity:
&1/2
32πe6
% % &
dW 1 2π hν
jνff ≡ = 2
Z ne ni exp − G¯ff (12)
dV dtdν (4π$0 )3 3me3/2 c3 3kT kT
where G¯ff is now the velocity averaged Gaunt factor. Its value is typically of order unity.
60
Points to note:
61
PHYS 4011 – HEA Lec. 7
In Lec. 5 (Sec. 5.1, eqn. 6), we found the expressions for the radiation field resulting from a
nonuniformly moving charge, viz.
! $
q R̂ 1" #
Erad (r, t) = × (R̂ − β) × β̇ (1)
4π"0 (1 − R̂ · β)3 R c
Here, R is the displacement from the retarded source point to the field point and βc is the
particle’s velocity evaluated at the retarded time. We used this equation in the nonrelativistic
limit (β # 1) to calculate the Poynting flux and then derive the Larmor formula (c.f. eqn. 9 in
Lec. 5) for the total power emitted by a nonrelativistic particle. How do we calculate the total
power emitted by a relativistic particle? We could follow the same procedure, keeping terms
involving β , but the derivation is complicated by the fact that the Poynting flux at the field point
where we observe the radiation is not the same as the rate at which the energy left the source
point, because the charge is moving. There is an easier way which takes advantage of the
Lorentz invariance of the total power. We then need to consider how the radiation is emitted in
the particle’s rest frame and how it is received in an observer’s rest frame.
62
Step 1. Consider an instantaneous rest frame K ! such that a particle has zero velocity at a
certain time and moves nonrelativistically for infinitessimally neighbouring times. We can then
use the Larmor formula to calculate the total power P ! emitted in K ! . We then need to how to
transform back to the rest frame K of an observer who measures P . The two references
frames have a relative speed βc. First note that P ! = dW ! /dt! , where dW ! is the energy
emitted in time dt! in frame K ! . These quantities transform as
dW = γdW ! , dt = γdt!
where γ = (1 − β 2 )1/2 is the Lorentz factor. Note that the Lorentz factors cancel out, i.e.
dW γdW ! dW !
P = = = = P! (2)
dt γdt! dt!
so the total emitted power is Lorentz invariant.
63
PHYS 4011 – HEA Lec. 7
µ0 q 2 |a! |2
Step 2. Now we have the identity P = P! = 6πc . In this expression, a! is the
3-acceleration in K ! and it is possible to relate this to the 4-acceleration, thereby allowing us to
express P in the more general covariant form (i.e. obeying special relativity). The
4-acceleration can be defined as
dv µ d2 xµ
aµ ≡ = = , a2 = aµ aµ = −a0 a0 + |a|2 (3)
dτ dτ 2
where v µ = γ(c, v) is the 4-velocity. Note that the 4-velocity and 4-acceleration are
orthogonal: aµ vµ = (dv µ /dτ )vµ = 12 d(v µ vµ )/dτ = 12 d(−c2 )/dτ = 0. Now in K ! , we
can write |a! |2 = a!2 + a!0 a!0 . But in the particle’s rest frame, v !µ = (c, 0) and since
a!µ vµ! = 0, we must have a!0 = 0. Thus, we can write
and
µ0 q 2 µ
P = a aµ (4)
6πc
which is now in a manifestly covariant form (i.e. can be evaluated in any frame).
64
Step 3. Since aµ aµ = |a|!2 , we can keep P in terms of the 3-acceleration and we can write
the acceleration in terms of components parallel and perpendicular to the particle’s velocity, i.e.
a!" and a!⊥ . The transformation properties of these components are
Thus, we have
µ0 q 2 ! 2 µ0 q 2 !2 µ0 q 2 4 2 2
P = |a | = (a" + a!2
⊥ ) = γ (γ a" + a2⊥ ) (5)
6πc 6πc 6πc
Clearly, the power emitted increases drastically as a particle becomes relativistic. This
expression can be written another way:
µ0 q 2 6 % 2 &
P = γ a − |β × a|2 relativistic Larmor formula (6)
6πc
We will use this later to derive an expression for power emitted in synchrotron radiation by
relativsitic electrons in a magnetic field.
65
PHYS 4011 – HEA Lec. 7
Whilst the total radiation power emitted by a relativistic particle is Lorentz invariant, its angular
distribution is not. Consider again the rest frame of the particle K ! . We want to find how the
emitted energy per solid angle transforms from K ! to an observer rest frame K . Suppose the
relative velocity v between these two frames is along the x-axis. Then the change in energy
transforms as
dW = γ(dW ! + vdp!x )
from the transformation properties of the 4-momentum pµ= (E/c, p) = m0 v µ . For
photons, we have |p| = E/c, so if we define an angle ϑ measured w.r.t. the x-axis, then
px = (E/c) cos ϑ, so
dW = γ(1 + β cos ϑ! )dW ! (7)
Now we want to find how much energy is radiated in the solid angle
dΩ! = sin ϑ! dϑ! dφ! = d cos ϑ! dφ! about ϑ! and we want to know how dW ! /dΩ!
transforms to give dW/dΩ, where dΩ = d cos ϑdφ.
66
!
Geometry for dipole emission from a particle instantaneously at rest in frame K .
67
PHYS 4011 – HEA Lec. 7
cos ϑ! + β
cos ϑ = (8)
1 + β cos ϑ!
Differentiating gives
d cos ϑ!
d cos ϑ =
γ 2 (1 + β cos ϑ! )2
The azimuthal angle is invariant, so dφ = dφ! . Thus,
If the emission is isotropic in the particle’s rest frame K ! , then in K , it will be peaked forward
in the direction of motion (i.e. ϑ → 0). We know that in K ! , we can use the dipole
approximation , which gives (c.f. Lec. 5)
dP ! µ0 q 2 ! 2 2 !
= a sin θ (11)
dΩ! 16π 2 c
where θ is the angle between the acceleration and the direction of emission. So in the
particle’s rest frame, the emission drops to zero in the direction of acceleration and peaks in
the direction perpendicular to it (i.e. the dipole torus pattern). Writing a! = a!" + a!⊥ and
using the transformations defined earlier yields
2 2 2
dP µ0 q 2 (γ a" + a⊥ ) 2 ! 3P sin2 θ !
= sin θ = (12)
dΩ 16π 2 c (1 − β cos ϑ)4 8πγ 4 (1 − β cos ϑ)4
Note the strong dependence on the factor 1 − β cos ϑ in the denominator. This term
dominates when ϑ → 0 and β → 1. In other words, the radiation is observed to be strong in
the forward direction with respect to the particle’s motion. This is referred to as relativistic
beaming. We still have to transform θ ! back to angles in K and this is difficult to do for the
general case. We can do it for some special cases.
69
PHYS 4011 – HEA Lec. 7
dP µ0 q 2 2 sin2 θ
= a (14)
dΩ 16π 2 c " (1 − β cos ϑ)6
which peaks in the forward direction at ϑ ∼ 1/γ .
70
2 ! sin2 ϑ cos2 φ
sin θ = 1 − 2 (15)
γ (1 − β cos ϑ)2
giving
µ0 q 2 2 sin2 ϑ cos2 φ
' (
dP 1
= a 1− 2 (16)
dΩ 16π 2 c ⊥ (1 − β cos ϑ)4 γ (1 − β cos ϑ)2
Again, this peaks in the forward direction near ϑ ∼ 1/γ , with a smaller peak at larger ϑ due to
the azimuthal dependence.
71
PHYS 4011 – HEA Lec. 7
1 + γ 2 ϑ2
1 − β cos ϑ ∼ (17)
2γ 2
Substituting this limit back into (12) we get for the parallel and perpendicular cases
72
PHYS 4011 – HEA Lec. 8
73
dp d
= (γmv) = qv × B (1)
dt dt
We now make the assumption that γ ≈ const. This is a valid assumption provided the
emitted radiation field does not have a back reaction on the particle’s motion (i.e. provided
d(γmc2 )/dt = qv · E ≈ 0). So γ ≈ const gives mγdv/dt = qv × B and we can
separate the velocity into components parallel and perpendicular to B:
dv! dv⊥ q
=0 , = v⊥ × B (2)
dt dt γm
Thus, v! = const and since the total |v| = const (because γ = const), then
|v⊥ | = const also. Thus, there is uniform circular motion in the plane normal to B and the
acceleration is perpendicular to the velocity in this plane.
74
PHYS 4011 – HEA Lec. 8
|q|v⊥ B
a⊥ = = Ωv⊥ (3)
γm
where Ω = |q|B/γm is the gyrofrequency
(or cyclotron frequency). We can substitute
this into the expression we found for total
power emitted by a relativistic particle with
Helical motion of an electron in a magnetic field
acceleration perpendicular to velocity, eqn.
B results from the combination of uniform motion
(5) in Lec. 7, viz. along B and circular motion perpendicular to B.
µ0 q 2 4 2 µ0 q 2 4 q 2 B 2 2 1 q4 2 2 2 2
P = γ a⊥ = γ v =⇒ P = γ β B sin α (4)
6πc 6πc γ 2 m2 ⊥ 6π#0 c m2
where α is the pitch angle between v and B. Note the dependence on m: synchrotron
emission is much less efficient for protons than for electrons.
75
For an isotropic distribution of velocities, it is necessary to average over all pitch angles for a
given speed β :
+1
β2 1 2 2
! !
2 2
&β⊥ ' = sin α dΩ = β 2 (1 − cos2 α) d cos α = β (5)
4π 2 −1 3
So the total power emitted by an electron, averaged over all pitch angles is
2 1 q4 2 2 2
P = γ β B (6)
3 6π#0 c m2
This is also sometimes expressed in terms of the Thomson cross-section,
8 2 e4
σT = πr0 = = 6.65 × 10−29 m2 (7)
3 6π#20 m2e c4
where r0 = e2 /(4π#0 me c) = 2.82 × 10−15 m is the classical electron radius, obtained by
equating the electrostatic potential energy, e2 /(4π#0 r0 ), with the rest mass energy me c2 . So
2 c 4
P = σT γ 2 β 2 B 2 = σT cγ 2 β 2 UB total synchrotron power (8)
3 µ0 3
where UB = B 2 /2µ0 is the magnetic energy density (i.e. energy per unit volume).
76
PHYS 4011 – HEA Lec. 8
In the ultrarelativistic limit, the electron velocity is close to the speed of light, so the electron
appears as though it is trying to catch up to the photons it produces.
An observer at rest will see a pulse of electromagnetic radiation E(t) confined to a time
interval δt∼ 1/f + T , where T = 2π/Ω = 2πγme /(eB) is the gyration period. Thus,
the spectrum will be spread over δω , Ω.
77
Consider the diagram here in which an observer’s line-of-sight intercepts the emission cones of
a relativistic electron. The observer will thus detect pulses of radiation from points 1 and 2
along the electron’s helical path. The times t1 and t2 at which the electron passes points 1
and 2 satisfy v(t2 − t1 ) = ∆s, where ∆s is the distance travelled by the electron along its
path. If a is the radius of curvature, then ∆s = a∆θ , where ∆θ in this diagram is just equal
to 2/γ , from the geometry, so ∆s = 2a/γ . But from the equation of motion, we also have
∆v evB sin α
)
∆t γm
78
PHYS 4011 – HEA Lec. 8
and so
∆ta ∼ (γ 3 Ω sin α)−1 (11)
So the time interval between pulses and the width of the individual pulses, E(t), are smaller
γ 3 . When we take the Fourier transform of the pulses,
than the gyration period by a factor ∼
we expect to get a broad spectrum up to some cutoff frequency ω ∼ 1/∆ta . In fact, in the
exact treatment to follow, we will use the following definition
3 3 3 eB 2
ωc ≡ γ Ω sin α = γ sin α critical frequency (12)
2 2 me
The spectrum should fall off sharply at frequencies above ωc . We can estimate that the power
per unit frequency emitted per electron is P (ω)∼ P/ωc F (ω/ωc ), where F (ω/ωc ) is a
dimensionless function that describes the correct behaviour of the spectrum near ωc . Using
eqn. (4) for P and ωc defined above, we have (for β ) 1)
1 e3 B
$ %
ω
P (ω) ∼ F sin α (13)
9π#0 c me ωc
80
PHYS 4011 – HEA Lec. 8
81
If the energy limits are sufficiently wide, then we can take x1 ) 0 and x2 ) ∞ and the
resulting integral is approximately constant. In that case, we have
This is a power law spectrum and the spectral index s of the emitted radiation spectrum is
directly related to the particle distribution index p:
1
s= (p − 1) spectral index (18)
2
Although this relation has been derived qualitatively here, the result is the same for the exact
treatment.
82
PHYS 4011 – HEA Lec. 9
Step 1 – set-up. The Poynting flux gives the power emitted per unit area (c.f. Lec. 5):
dW 1
= |Erad (t)|2 (1)
dtdA µ0 c
and the radiation field for a moving electron is that given by eqn. (6) in Lec. 5, viz
! $
e R̂ 1" #
Erad (r, t) = × (R̂ − β) × β̇ (2)
4π"0 (1 − R̂ · β)3 R c
where R̂ is the propagation direction of the radiation to the observer. Note that the quantities
on the RHS are evaluated at the retarded time t! .
83
dW
= 2c"0 |E(ω)|2 (5)
dAdω
84
PHYS 4011 – HEA Lec. 9
Step 2 – Fourier transform of the retarded field. We need to solve the integral on the RHS of
+∞
1 e R̂ × [(R̂ − β) × β̇]
%
RE(ω) = eiωt dt (7)
(2π)1/2 4πc"0 −∞ (1 − R̂ · β)3
We first change the integration variable from t to t! , using the definition of retarded time
t! = t − R(t! )/c and R = |r − r0 | to give
dt!
dt = = (1 − R̂ · β) dt! (8)
(∂t! /∂t)
So now we have
+∞
1 e R̂ × [(R̂ − β) × β̇]
%
RE(ω) = eiωt dt! (9)
(2π)1/2 4πc"0 −∞ (1 − R̂ · β)2
Next we express eiωt = exp[iω(t! + R/c)] in terms of t! only. We assume the radiation is
being observed far enough away from the source that r(t) $ r0 (t! ), so that R # r .
85
Then we expand r to first order in r0 , which gives R(t! ) # |r| − R̂ · r0 . Now we have
+∞
1 e R̂ × [(R̂ − β) × β̇]
% " ' (#
RE(ω) = 1/2
exp iω t! − R̂ · r0 (t! )/c dt!
(2π) 4πc"0 −∞ (1 − R̂ · β)2
(10)
Then we use the identity
dW e2 ω 2 )) +∞
)% " ' (# ))2
= R̂ × (R̂ × β) exp iω t! − R̂ · r0 (t! )/c dt! )) (11)
dωdΩ 16π 3 "0 c ) −∞
This is now in a form that can be integrated.
86
PHYS 4011 – HEA Lec. 9
Step 3 – evaluation of integral. To simplify the integration, we need to simplify the triple cross
product. Consider the diagram below.
An electron moves along an orbital trajectory
with radius of curvature a. The coordinate
system is set up such that the electron is trav-
elling in the x − y plane and passes through
the origin at retarded time t!
= 0 with an in-
stantaneous velocity in the x-direction. The
unit vector e⊥ is along the y axis and e% =
R̂ × e⊥ . Thus, e% and e⊥ define a plane
perpendicular to an observer’s line of sight de-
fined by the direction R̂. This is the plane of
propagation, defined by the triple cross prod-
uct R̂ × (R̂ × β) in the integral.
The magnetic field B must also be in the plane containing R̂ and β , so e% and e⊥ define
directions parallel and perpendicular to the projection of the magnetic field on the plane of
propagation.
87
dW dW% dW⊥
= + (14)
dωdΩ dωdΩ dωdΩ
and where, defining ϑ2γ = 1 + γ 2 ϑ2 , we have
+- )2
dW% e2 ω 2 ϑ2 )) c2 γ 2 t!3
)% , *
iω 2 ! !)
)
= exp ϑ γ t + dt
dωdΩ 16π 3 "0 c ) 2γ 2 3a2 )
+- )2
e2 ω 2 )) ct! c2 γ 2 t!3
)% , *
dW⊥ iω 2 ! !)
)
= exp ϑ γ t + dt (15)
dωdΩ 16π 3 "0 c ) a 2γ 2 3a2 )
88
PHYS 4011 – HEA Lec. 9
ct! ωaϑ3γ
y≡γ , η =≡
aϑγ 3cγ 3
which gives
+2 )% +∞ +- )2
dW% e2 ω 2 ϑ2
* , *
aϑγ) 3 1 3
)
= ) exp iη y + y dy )
dωdΩ 16π 3 "0 c γc)
−∞ 2 3 )
. /2 )% +- )2
e2 ω 2 ϑ2 aϑ2γ ) +∞
, *
dW⊥ 3 1 3
)
= ) y exp iη y + y dy ) (16)
dωdΩ 16π 3 "0 c γ 2 c )
−∞ 2 3 )
The integrals can be expressed in terms of the modified Bessel functions of 1/3 and 2/3
order:
+2
dW% e2 ω 2 ϑ2
*
aϑγ 2
= K1/3 (η)
dωdΩ 16π 3 "0 c γc
. /2
dW⊥ e2 ω 2 ϑ2 aϑ2γ 2
= K2/3 (η) (17)
dωdΩ 16π 3 "0 c γ2c
89
We next integrate over solid angle to give the energy per frequency emitted by an electron per
orbit in the plane of propagation. During one orbit, the emission is almost completely confined
to within an angle 1/γ around a cone of half-angle α (the pitch angle). So we use
dΩ # 2π sin αdϑ. Thus,
+∞
dW% e2 ω 2 a2 sin α
%
# ϑ2γ ϑ2 K1/3
2
(η)dϑ
dω 6π 2 "0 c3 γ 2 −∞
2 2 2 +∞
dW⊥ e ω a sin α
%
= ϑ4γ K2/3
2
(η)dϑ (18)
dω 6π 2 "0 c3 γ 4 −∞
3 eB 2
and x = ω/ωc , where ωc = 2 me γ sin α is the critical frequency (see Lec. 8). To convert
this to power per unit frequency, we divide by the orbital period T = 2π/Ω = 2πγme /eB ,
which gives
√ 3
3e B sin α
P% (ω) = [F (x) − G(x)]
16π 2 "0 me c
√ 3
3e B sin α
P⊥ (ω) = [F (x) + G(x)] (21)
16π 2 "0 me c
These are the components of the single-electron synchrotron power per unit frequency
corresponding to polarisation modes parallel and perpendicular to B. The total synchrotron
power per unit frequency is
√
3e3 B sin α
P (ω) = P% (ω) + P⊥ (ω) = F (x) (22)
8π 2 "0 me c
single electron synchrotron power spectrum
91
The functions F (x) and G(x) that appear in the synchrotron power spectrum are plotted
below in linear and logarithmic scales.
Both functions have similar shapes and reach similar asympototic values at large x, but
G(x) < F (x) for x < 1. The asympototic behaviour of the functions goes as
F (x), G(x) ∼ x1/2 e−x , x $ 1 F (x), G(x) ∼ x1/3 , x ( 1 (23)
for the number of electrons with energy γme c2 . The total number of electrons per unit volume
& γ2
is Ne = γ1
Ne (γ) dγ . Note that the electron pitch angles α will in general also be spread
around a direction kα defined by a characteristic angle α0 between R̂ and B. We assume
the electron distribution is isotropic in pitch angle, so that Ne (γ) is independent of kα .
The synchrotron volume emissivity for a nonthermal distribution of relativistic electrons thus
has the following form:
% γ2 % γ2
jνsyn = 2πKe P (ω) γ −p
dγ ∝ F (x)γ −p dγ (25)
γ1 γ1
&
P (ω) = dW/(dωdt) = dW/(2πdνdt) = jν dV /2π . We now change the integration
variable from γ to x = ω/ωc using the relations
* +−1/2 * +−1/2
3 eBx sin α 1 3 eB sin α
γ= , dγ = − x−3/2 dx (26)
2 ωme 2 2 ωme
So we now have an integral of the form
% x1
jνsyn ∝ x(p−3)/2 F (x) dx (27)
x2
To evaluate this integral analytically, the following approximation is made. Consider the
relations
ω ω
x1 = , x2 =
ωc (γ1 ) ωc (γ2 )
If γ1
( γ2 , then ωc (γ1 ) ( ωc (γ2 ) and we can have a wide range of frequencies satisfying
ω $ ωc (γ1 ) so that x1 → ∞. Similarly, we can have ω ( ωc (γ2 ), so that x2 → 0. Thus,
94
PHYS 4011 – HEA Lec. 9
p 19 p 1
jνsyn = 3p/2 2−(p+7)/2 π −(p+3)/2 Γ( + )Γ( − )(p + 1)−1 (28)
4 12 4 12
(p+1)/2
e2
* +
eB
Ke sin α0 ν −(p−1)/2 synchrotron emissivity
"0 c me
This is clearly a power-law spectrum, with spectral index
1
α= (p − 1) (29)
2
If the magnetic field B does not have a fixed direction (e.g. randomly oriented), then we need
95
96
PHYS 4011 – HEA Lec. 9
The final and most often used expression for the synchrotron emissivity is the following for the
randomly oriented magnetic field case:
Γ( 5+p p
4 )Γ( 4 +
19 p 1
12 )Γ( 4 − 12 )
jνsyn = 3p/2 2−(p+13)/2 π −(p+2)/2
Γ( 7+p
4 )
+(p+1)/2
e2
*
eB −(p−1)/2
Ke ν (31)
"0 c me
synchrotron emissivity (pitch angle averaged)
97
Compton scattering is the scattering of photons off electrons. For low photon energies, it
reduces to the classical case of Thomson scattering. For relativistic electrons, lower energy
photons can be efficiently upscattered to energies reaching X-ray and γ -ray wavelengths. The
photon upscattering process is referred to as Comptonisation. When referring to cooling of the
electrons, the radiation process is called inverse Compton scattering. The emission (scattered)
spectrum can be calculated analytically for single scatterings only. For multiple scatterings,
numerical simulations are usually necessary.
99
100
PHYS 4011 – HEA Lec. 10
these in, rearranging and using k̂i · k̂f = cos Θ gives the following expression for the energy
of the scattered photon:
ε
ε1 = ε (3)
1+ me c2 (1 − cos Θ)
101
∆λ = λ1 − λ = λC (1 − cos Θ) (4)
where
h
λC ≡ # 0.0243 Å Compton wavelength (5)
me c
Thus, the wavelength change is of order λC . For long wavelengths (λ $ λC ) or equivalently,
ε % me c2 , the scattering is approximately elastic (i.e. ε1 # ε). This is the Thomson regime.
102
PHYS 4011 – HEA Lec. 10
In addition to the effects of photon momentum, quantum corrections also modify the cross
section for Compton scattering. The exact expression for the differential cross section for
Compton scattering is derived from quantum electrodynamics and is known as the
Klein-Nishina formula:
1 ε2
# $
dσKN ε ε1
= r02 12 + − sin2 Θ (6)
dΩ 2 ε ε1 ε
where r0 = e2 /(4π'0 me c) = 2.82 × 10−15 m is the classical electron radius (defined in
Lec. 8). This reduces to the classical differential Thomson cross section in the limit ε1 ∼ ε,
viz. dσT /dΩ = 12 r02 (1 + cos2 Θ). The total cross section is obtained by integrating over
% +1
solid angle, σKN = 2π −1 (dσKN /dΩ) d cos Θ:
& ' ( )
3 1 + x 2x(1 + x) 1 1 + 3x
σKN = σT − ln(1 + 2x) + ln(1 + 2x) −
4 x3 1 + 2x 2x (1 + 2x)3
(7)
2
where x ≡ hν/me c .
103
The overall effect of σKN is to reduce the scattering cross section relative to σT at high photon
energies. Thus, Compton scattering becomes less efficient at high energies. The decline is
shown in the plot below.
104
PHYS 4011 – HEA Lec. 10
In general, electrons will not be at rest, but will be moving, sometimes with relativistic
velocities. Whenever a moving electron has energy greater than that of an incident photon, the
energy transfer is from electron to photon. This is inverse Compton scattering. The results for
scattering by a stationary electron are extended to a moving electron using a Lorentz
transformation. Let K be the observer’s frame and K " be the rest frame of an electron. The
= (1 − β 2 )−1/2 . A scattering event as seen
relative velocity βc defines the Lorentz factor γ
in each frame is shown in the figure below. In K , the electron’s velocity is in the x direction
and all angles in both frames are measured from this axis.
!
observer’s frame K electron rest frame K
105
In K " , all the previous formulas for scattering from stationary electrons are valid. Transforming
the photon’s initial energy into K " :
Thus, in transforming to the electron rest frame, the photon picks up a factor γ and in
transforming back to the lab frame, it picks up an additional factor γ . Hence, the photon energy
can increase by a factor γ 2 in the lab frame, implying that Compton scattering by relativistic
electrons can be quite efficient.
A maximum gain of ∼ γ 2 is only possible for scatterings that are in the Thomson regime in the
rest frame (i.e. ε"1 # ε" ) and which have Θ, Θ"1 > ∼ π/2. The condition for Thomson scattering
in the rest frame is
ε" % me c2 =⇒ γhν % me c2 (10)
106
PHYS 4011 – HEA Lec. 10
dE1"
*
= cσT ε"1 n" (ε" ) dε" (11)
dt"
Now we know that the emitted power is an invariant. Another invariant is the quantity
n(ε)dε/ε. So
dE1 ndε
* *
= cσT ε n dε = cσT ε"2
" " "
(12)
dt ε
"
Now we substitute ε = εγ(1 − β cos Θ) from eqn. (8) to get
dE1
*
2
= cσT γ (1 − β cos Θ)2 εn dε (13)
dt
107
which now only contains quantities in frame K . For an isotropic distribution of photons, we
have
1
)(1 − β cos Θ)2 * = 1 + β 2
3
giving
# $
dE1 2 1 2
= cσT γ 1 + β Uγ (14)
dt 3
%
where Uγ = εndε is the initial photon energy density. Now dE1 /dt is the rate at which the
electron loses energy. The nett power converted into increased radiation is this minus the rate
at which the initial photon energy distribution decreases, dε/dt
= σT cUγ . So
' # $ (
dErad dE1 dε 2 1 2
= − = cσT Uph γ 1 + β − 1 (15)
dt dt dt 3
which gives the following for the inverse Compton power for a single electron:
4
Pic = σT cγ 2 β 2 Uγ inverse Compton power (16)
3
For a nonthermal power law distribution of electrons N (γ) = Ke γ −p , we can obtain the total
power per unit volume from
* γ2
Pic,tot = Pic N (γ)dγ
γ1
4
Pic,tot = σT cUγ Ke (3 − p)−1 (γ23−p − γ13−p ) nonthermal power-law electrons (17)
3
For a thermal distribution of electrons, γ = 1 and )β 2 * = 3kTe /me c2 . The total power
needs to be derived from the single electron power in the more general case where energy
transfer in the electron rest frame is not neglected. The result is
4kTe
Pic,tot = σT cUγ Ne thermal electrons (18)
me c2
109
110
PHYS 4011 – HEA Lec. 10
3 p2 + 4p + 11
*
jεic1 = 2p−2 ε(p−1)/2 n(ε) dε
−(p−1)/2
σT cε1 Ke (20)
π (p + 1)(p + 3)2 (p + 5)
Thus, inverse Compton scattering also predicts a power law spectrum with a spectral index
1
α= (p − 1) (21)
2
identical to the case of synchrotron emission. The power law spectrum is independent of the
incident photon distribution.
111
The above derivation implies that if the incident photon distribution is a blackbody spectrum,
the resulting spectrum after a single scattering by nonthermal electrons should be a power law.
For a blackbody, we have
8π 2 + , ε - .−1
n(ε) = ε exp − 1 (22)
(hc)3 kT
Inserting this into the expression jε1 above, and solving the integrals gives
σT
jεic,bb fbb (p)(kT )(p+5)/2 Ke ε1
−(p−1)/2
= 3 2
(23)
1
h c
where
2p+1 (p2 + 4p + 11)
# $ # $
p+5 p+5
fbb (p) = 3 γ ζ
(p + 1)(p + 3)2 (p + 5) 2 2
where ζ is the Riemann zeta function.
112
PHYS 4011 – HEA Lec. 10
Synchrotron Self-Comptonisation
A particularly interesting case of inverse Compton scattering is that in which the seed photons
are synchrotron photons emitted by the scattering electrons. In this case, the incident photon
spectrum is the synchrotron power law spectrum, which can be written as
# $−(p−1)/2
Uγ (ε0 ) ε
n(ε) = , εmin <
∼ εmax (24)
ε0 ε0
where ε0 is some fiducial seed photon energy. The solution for the synchrotron self-Compton
volume emissivity is
# $# $−(p−1)/2
εmax ν1
jνssc = f (p)σT cKe Uγν0 ln (25)
1
εmin ν0
where the the relation jν1 = hjε1 has been used and where
3 p−2 p2 + 4p + 11
f (p) = 2
π (p + 1)(p + 3)2 (p + 5)
The term ln(εmax /εmin ) is known as the Compton logarithm.
113
The mean number of scatterings is determined by the optical depth, τ = σNe r , where r is
the size of the scattering region. A value of τ ∼ 1 means that on average, a photon will scatter
once before escaping the region. Specifically, we have
y<
∼1 power law spectrum, with exponential cut-off (28)
114
PHYS 4011 – HEA Lec. 10
For y <
∼ 1, it is possible to obtain a power law scattered spectrum, even if the scattering
electrons have a thermal distribution. For a thermal electron distribution, the Compton y
parameter is defined as
# $
4kTe 4kTe 2
y= 1+ max(τT , τT ) (29)
me c2 me c2
y parameter for thermal Comptonisation (30)
where τT = σT Ne r is the Thomson optical depth of the scattering region of size r . In the
saturated Comptonisation limit (y $ 1), the incident photon spectrum is completely distorted
beyond recognition. The resulting spectrum approaches a similar distribution to the scattering
electrons, implying that the photons and electrons come into thermal equilibrium. An incident
nonthermal (i.e. power law) photon spectrum, for example, will become thermalised by the
scatterings and approach a blackbody spectrum at the temperature of the scattering electrons,
so the spectrum will peak at hν1 # 2.8kTe .
115
Some example spectra of multiple Compton scatterings calculated from Monte Carlo
simulations (see e.g. Sunyaev & Titarchuk, 1980, Astron. Astrophys., 86, 121.):
116