Radiation Processes: 1.1 Review of Basic Definitions

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Chapter 1

Radiation Processes

Our knowledge of astronomy is largely based on observations of electromagnetic


radiation. It is therefore essential to understand the radiation generation and
propagation processes, such that we can infer the nature and environment of
celestial objects from observations.

In this chapter we first review some basics of radiative transfer and then in-
troduce a few common radiation mechanisms related to this course, including
blackbody, synchrotron, and inverse-Compton. Note that there are some
other radiation processes not covered by this course. For example, bremsstrahlung
happens when a charged particle being scattered by another charged particle.
Cherenkov radiation occurs when a particle moves faster than the speed of
light in a medium.

1.1 Review of basic definitions

Specific intensity Iν , is defined as the energy carried by photons pass through


a small area dA perpendicular to some direction ~n, pointing to similar directions
within a small solid angle dΩ, between frequency ν and ν + dν, and time dt, i.e.
dEν
Iν ≡ . (1.1)
dA dt dΩ dν
It has units of erg/s/cm2 /sr/Hz.

The monochromatic emission coefficient jν is defined as the energy emitted

1
CHAPTER 1. RADIATION PROCESSES 2

per unit time per unit solid angle per unit volume, i.e.
dE
jν ≡ (1.2)
dV dΩdt dν
and it has units of erg/cm3 /s/sr/Hz. When a beam of cross section dA travels
though a emission region for a distance ds, the volume it covers is dV =dAds.
The energy it gains is
dIν = jν ds . (1.3)

The absorption coefficient αν is defined as the change in the beam intensity


after traveling for a distance ds:

dIν = −αν Iν ds . (1.4)

αν is in units of cm−1 . Putting Eqs. (1.3) and (1.4) together, the general form
of the radiative transfer equation is
dIν
= −αν Iν + jν . (1.5)
ds
For pure emission, αν = 0, the solution is
Z s
Iν (s) = Iν (s0 ) + jν (s0 ) ds0 . (1.6)
s0

For pure absorption, jν = 0, the solution is


Rs
− αν (s0 ) ds0
Iν (s) = Iν (s0 )e s0
. (1.7)

The integral in the exponent is called the optical depth


Z s
τν ≡ αν (s0 ) ds0 . (1.8)
s0

In the following sections, we will introduce some common emission mechanisms


and absorption processes.

1.2 Blackbody radiation

Here we only give a brief review of basic properties of blackbody radiation, as


the details are covered in other courses. Photons in thermal equilibrium obey
Bose-Einstein statistics, such that the number density between frequency ν
and ν + ∆ν is
8πν 2 dν
n(ν) dν = 3 . (1.9)
c exp(hν/kT ) − 1
CHAPTER 1. RADIATION PROCESSES 3

This is also known as the blackbody spectrum.

A simple way to understand this is that the energy of having s photons each
with momentum p~ equals to spc. As photons are bosonic, s can take on any
natural number. The probability P (s) that there are exactly s photons each
with momentum p~ follows the constraint
P (s + 1)
= e−pc/kT = e−hν/kT , (1.10)
P (s)
where ν is the frequency of the photon. Consequently,

e−shν/kT
P (s) = P+∞ −s0 hν/kT = e−shν/kT (1 − e−hν/kT ) (1.11)
s0 =0 e

provided that p~ is an allowed momentum of a photon. Thus, the expected


number of photons with an allowed momentum p~ is given by
+∞
X 2
hNγ (~p)i = 2 sP (s) = . (1.12)
s=0 ehν/kT −1

(Note that the 2 above reflects the fact that a photon has two possible polar-
izations.) As a result, the total expected number of photons equals
1 Z Z
hNγ i = 3
hNγ (~x)i dV d3 p~
h Z
1 Z
2
= 3 dV hν/kT
d3 p~
h e −1
Z
8πν 2
= V dν . (1.13)
c3 (ehν/kT − 1)

For a photon of momentum p~ hitting a wall and then reflected back elastically,
the magnitude of the change in momentum equals 2p cos θ = 2hν cos θ/c where
θ is the angle between p~ and the normal of the wall. The radiation pressure is
due simply to the momentum transfer to the wall per unit time per unit surface
area. That is,
!2 !
1 Z +∞ Z π/2 Z 2π 2hν cos θ 2 hν hν
Prad = 3 c cos θ sin θ dφdθd
h 0 0 0 c ehν/kT − 1 c c
Z +∞

= n(ν) dν
0 3
4σ 4
= T , (1.14)
3c
where σ = 2π 5 k 4 /15c2 h3 is the Stefan’s constant.
CHAPTER 1. RADIATION PROCESSES 4

Finally, note that the energy density of a photon gas at temperature T is given
by Z +∞
4σ 4
urad = hνn(ν)dν = T . (1.15)
0 c
Hence, urad = 3Prad .

Exercise:
1. For a photon gas with volume V and temperature T , what is the internal
energy U ?
2. Given that the entropy S of the gas is related to U by dU = T dS and S
remains a constant for an adiabatic process, show that a photon gas has
an adiabatic index γ = 4/3.

1.3 Synchrotron radiation

Charged particles radiate when they are being accelerated. When the accelera-
tion is caused by magnetic field, the radiation is called magnetobremsstrahlung.
It has different names according to the particle energy. It is called gyro radi-
ation for non-relativistic particles, cyclotron radiation for mildly relativistic
particles, and synchrotron radiation for ultra-relativistic particles. The last
type is called curvature radiation when the magnetic field is extremely strong
(e.g., near neutron stars) such that the charged particles are forced to move
along the magnetic field lines. In this section we focus on synchrotron radiation
because it is a common process in astronomical objects.

For non-relativistic particles, the gyro frequency and angular frequency are given
by
qB sin α qB sin α
νG ≡ and ωG ≡ , (1.16)
2πmc mc
respectively, where α is the pitch angle between the velocity and the magnetic
field. For relativistic particles, the gyration frequency ωB is lower because of
relativistic mass,
ωG
ωB = . (1.17)
γ
CHAPTER 1. RADIATION PROCESSES 5

Figure 1.1: Emission cones at various points of an accelerated particle’s trajec-


tory (Rybicki & Lightman, 1979).

This is however not the actual spectral frequency that would be observed as we
will discuss below.

The emission from a relativistic particle is highly beamed, with a half-opening


angle ∼1/γ. The emission is therefore only visible when the particle is moving
nearly towards the observer, as shown in Figure 1.1. For radiation emitted at
positions 1 and 2, the difference in arrival time is
∆s s − ∆s s ∆s v
     
∆tp = + − = 1− , (1.18)
v c c v c
where s is the distance from position 1 to the observer, ∆s is the distance
between positions 1 and 2, and v is the particle velocity. Note that

∆s = a∆θ,

where a is the gyro radius and ∆θ = 2/γ as shown in Figure 1.1. Also v = aωB
gives
∆s/v = ∆θ/ωB .
Combining the above equations and using the approximation that (1 − v/c) ≈
1/(2γ 2 ) for γ  1, one can derive
2 1 1
∆tp = 2
= 3 . (1.19)
γωB 2γ γ ωB

Applying Fourier transform to a pulse train of interval ∆tp 1 , it can be shown


that the shape of the synchrotron spectrum is relatively flat up to the critical
1
See https://www.cv.nrao.edu/~sransom/web/Ch5.html for the details.
CHAPTER 1. RADIATION PROCESSES 6

frequency νc = 1/∆tp and then falls off rapidly. The spectral shape is plotted
in Fig. 1.2. From a more rigorous derivation, the critical frequency has an
additional factor of 3/2
3
νc = γ 3 νB . (1.20)
2
and the full synchrotron spectrum is
√ 3
3e B sin α ν Z ∞
 
Psyn (ν, E) = K5/3 (η)dη, (1.21)
me c2 νc ν/νc
where K5/3 is a modified Bessel function. The critical frequency is usually taken
to be the characteristic frequency of the synchrotron radiation.

R∞
Figure 1.2: Synchrotron radiation spectrum F (x) ≡ x x K5/3 (η)dη, where
x = ν/νc .

In all astrophysical situations, synchrotron radiation is emitted by a collection


of electrons instead of a single electron. Relativistic electrons (sometimes called
non-thermal electrons) usually follow a power-law energy distribution

n(E)dE ∝ E −p dE. (1.22)

The overall spectrum is then


Z
S(ν)dν = n(E)Psyn (ν, E)dE
ν
Z  
−p
∝ E Psyn dE
E2
∝ ν −(p−1)/2 . (1.23)
CHAPTER 1. RADIATION PROCESSES 7

In other words, the observed spectrum is also a power law, and the spectral
index α, which is defined as S(ν) ∝ ν −α , is related to the particle distribution
by α = (p−1)/2. Observation of the synchrotron spectrum (e.g., from supernova
remnants or AGN), can therefore reveal the underlying particle distribution and
hence the acceleration mechanism.

To calculate the total synchrotron power radiated by a single particle, one can
use the Larmor formula
2q 2 γ 6 h 2 i
P = (β̇ − (β × β̇)2 , (1.24)
3c
where β = v/c. For the non-relativistic version, just assume γ = 1 and β = 0.
For synchrotron radiation, one can derive the power to be

P = 2σT β 2 γ 2 cUB sin2 α,


4
hP i = σT β 2 γ 2 cUB , (1.25)
3
where σT = 8πe4 /3c4 m2e is the Thomson cross section and UB = B 2 /8π is the
magnetic energy density. The characteristic cooling time is defined as tcool =
E/P ∝ 1/E. Therefore, more energetic particles have shorter cooling time, i.e.
cool faster.

Synchrotron radiation is highly linearly polarized and slightly circularly po-


larized. For electrons with a power-law energy distribution, the linear polariza-
tion fraction can be shown to be
1+p
PF = . (1.26)
7/3 + p

The index p is usually larger than 1 for most astronomical objects, such that
the local polarization fraction is usually at least 60%. The polarization angle
is always perpendicular to the magnetic field, so the magnetic field lines can be
probed by observing the polarization direction. This is how the magnetic field
around the supermassive black hole M87 was revealed.

1.4 Inverse-Compton scattering

When a photon collides with a standing electron, it changes direction and


loses energy. This process is called Compton scattering. When a low-energy
photon (e.g., radio photon) collides with a relativistic electron, it inversely gains
energy and this is called inverse-Compton (IC) scattering. IC scattering
CHAPTER 1. RADIATION PROCESSES 8

can easily accelerate radio photons to gamma-ray photons and is very important
in explaining the gamma-ray emission observed in the Universe.

Compton scattering
We first review the derivation of Compton scattering. We use natural units
where c = 1. Assuming the initial 4-momentum of the photon and the elec-
tron are p1 = (E1 , E1 , 0, 0) and p2 = (me , 0, 0, 0), respectively. After collision,
they become p3 = (E2 , E2 cos θ, E2 sin θ, 0), and p4 , respectively, where θ is the
scattering angle. By conservation of 4-momentum,

p1 + p2 = p 3 + p4
(p1 + p2 − p3 )2 = p24
m2e + 2me E1 − 2me E2 − 2E1 E2 (1 − cos θ) = m2e
1 1 1
− = (1 − cos θ). (1.27)
E2 E1 me
The result is more commonly written in terms of wavelength λ = hc/E.
h
λ2 − λ1 = (1 − cos θ). (1.28)
me c
For unpolarized photon, the scattering cross section is given by the Klein-
Nishina formula, which is derived from quantum electrodynamics, and it
reduces to the Thomson cross section for non-relativistic particle.

Figure 1.3: Scattering geometries in the observer’s frame S and in the electron
rest frame S 0 .

Inverse-Compton scattering
We now consider the case where the electron is moving. The symbols are defined
in Figure 1.3. In the lab frame S, the electron moves in speed β = v/c in the
x-direction. In the electron’s rest frame S 0 , the problem becomes the same as
CHAPTER 1. RADIATION PROCESSES 9

the previous one. For low energy incident photons (E1 , E10  me c2 ), we rewrite
Eq. 1.27 as

E10
" #
E20
≈ E101− (1 − cos Θ) , (1.29)
me c2
cos Θ = cos θ10 cos θ20 + sin θ10 sin θ20 cos(φ01 − φ02 ), (1.30)

where φ01 and φ02 are the azimuthal angles of the scattered photon and incident
photon in the rest frame. Using Lorentz transformation (or Doppler shift), we
have

E10 = E1 γ(1 − β cos θ1 ), E2 = E20 γ(1 + β cos θ20 ). (1.31)

Combining the equations, the energy of the scattered photon is


E1
 
E2 = E1 γ 2 (1 − β cos θ1 )(1 + β cos θ20 ) 1 − (1 − cos Θ) . (1.32)
me c2
Neglecting the angular dependence, the energy of a photon in general increased
by a factor of γ 2 . For an electron with γ = 1000, a radio photon (e.g., 10 GHz)
can be accelerated to X-ray range (1016 Hz ∼ 40 eV). The maximum energy is
achieved when θ1 = π and θ2 = 0, representing a head-on collision in which
the photon is sent back along its original path. The maximum energy of the
scattered photon is

E2,max ≈ E1 γ 2 (1 + β)2 ≈ 4γ 2 E1 . (1.33)

IC emission power
The cosmic microwave background (CMB) provides a wealth of low energy
photons. Wherever relativistic plasma is present, IC scattering of CMB would
make it glows in X-rays or γ-rays. Here we calculate the power of this radiation
process. Let the energy density of background radiation be urad in the observer
frame. In this frame, the background radiation is isotropic. In the electron rest
frame, the radiation becomes anisotropic for two reasons, Doppler shift and
change of photons collision rate. The Doppler effect is given by

E10 = γE1 (1 + β cos θ). (1.34)

The collision rate follows a similar argument as Doppler shift, and therefore
gives rise to the same factor. Combing the two effects, the energy density of
the radiation in electron rest frame is
Z π
21 4 1
 
u0rad = urad γ (1 + β cos θ) sin θdθ = urad γ 2 −
2
. (1.35)
0 2 3 4
CHAPTER 1. RADIATION PROCESSES 10

Derived from Thomson scattering, the energy loss rate of the electron is given
by !0
dE
− = σT cu0rad . (1.36)
dt
Since radiation loss is Lorentz invariant, the energy loss rate is the same in all
reference frame, i.e.
!0
dE dE 4 1
 
= = − σT curad γ 2 − . (1.37)
dt dt 3 4
This is the energy radiated by the electron. The electron actually absorbs some
energy from the background radiation before re-radiation, so the net energy
loss of the electron should be subtracted by the absorbed background radiation,
which is
dE 4 1
− = σT curad (γ 2 − ) − σT curad
dt 3 4
4
= σT curad (γ 2 − 1)
3
4
= σT curad β 2 γ 2 . (1.38)
3
The result also tells the average energy gain of a photon
4
hE2 i ≈ γ 2 E1 . (1.39)
3

This is remarkable to the synchrotron radiation power in Eq. 1.25


dE 4
− = σT cuB β 2 γ 2 .
dt (syn) 3
The reason for this is that the energy loss rate depends upon the electric field
which accelerates the electron in its rest frame and it does not matter what
the origin of that field is. In the case of synchrotron radiation, the electric
~ field due to motion of the electron through the magnetic
field is the (~v × B)
field whereas, in the case of inverse Compton scattering, it is the sum of the
electric fields of the electromagnetic waves incident upon the electron. In the
latter case, the sum of the squares of the electric field strengths appears in the
formulae for incoherent radiation and so the energies of the waves add linearly.
Another way to understand this is that for synchrotron radiation, an electron
interacts with virtual photons generated by the magnetic field. It is therefore
not surprising to see that the two formula have very form.

Suppose synchrotron radiation and IC scattering is produced by the same source


of electrons, their relative luminosity is
Lsyn uB
= . (1.40)
LIC urad
CHAPTER 1. RADIATION PROCESSES 11

Synchrotron radiation and IC in general occupies a different frequency range, so


observationally it is possible to separate their luminosity. If we further assume
that CMB is the only background photon field, the magnetic field strength of the
plasma can be found. Estimating the magnetic field strength of astrophysical
plasma is a generally difficult task and this is one of the few available methods.
(Do you know any other methods?)

You might also like