You are on page 1of 150

Radiative Processes in Astrophysics

Jean-Pierre Macquart

Lecture notes based on: Radiative Process in Astrophysics, Rybicki & Lightman, Wiley Electromagnetic Processes in Dispersive Media, Melrose & McPhedran, Cambridge University Press

Contents
1 Fundamentals of Radiative Transfer 1.1 The Electromagnetic Spectrum . . . . . . . . . . . . . . 1.2 Geometric considerations . . . . . . . . . . . . . . . . . 1.2.1 Specic intensity . . . . . . . . . . . . . . . . . . 1.2.2 Net ux, momentum ux and radiation pressure 1.2.3 Radiative energy density . . . . . . . . . . . . . . 1.3 Radiative Transfer . . . . . . . . . . . . . . . . . . . . . 1.3.1 Emission . . . . . . . . . . . . . . . . . . . . . . 1.3.2 Absorption . . . . . . . . . . . . . . . . . . . . . 1.3.3 The equation of radiative transfer . . . . . . . . 1.4 Thermal radiation . . . . . . . . . . . . . . . . . . . . . 1.4.1 Thermodynamics of blackbody radiation . . . . . 1.4.2 The Plank Spectrum . . . . . . . . . . . . . . . . 1.4.3 Temperatures associated with emission . . . . . . 1.5 Einstein coecients . . . . . . . . . . . . . . . . . . . . . 1.5.1 Maser emission . . . . . . . . . . . . . . . . . . . 1.5.2 Line proles . . . . . . . . . . . . . . . . . . . . . 1.6 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9 9 10 10 12 14 15 15 16 17 19 20 22 23 24 27 27 28 30 30 31 32 33 34 35 37 38 38

2 Review of Electromagnetic Theory 2.1 Maxwells Equations and the Poynting Vector . . . . . . . . . . . 2.1.1 Continuity equations . . . . . . . . . . . . . . . . . . . . . 2.2 Electromagnetic Potentials . . . . . . . . . . . . . . . . . . . . . 2.2.1 Gauge conditions . . . . . . . . . . . . . . . . . . . . . . . 2.3 Plane Electromagnetic Waves . . . . . . . . . . . . . . . . . . . . 2.4 Polarization: The Mutal Coherence Matrix and Stokes Parameters 2.5 Fourier relations for electromagnetic waves . . . . . . . . . . . . . 2.6 The Geometrical Optics Limit . . . . . . . . . . . . . . . . . . . . 2.7 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

3 Electromagnetic Radiation from Moving Charges 40 3.1 Li enard-Wiechart Potentials . . . . . . . . . . . . . . . . . . . . . 40 3.1.1 The Lienard-Wierchert potentials . . . . . . . . . . . . . . 40 3.1.2 The Poynting Vector . . . . . . . . . . . . . . . . . . . . . 41

3.2 3.3 3.4 3.5 3.6 3.7

3.1.3 Larmor Formula . . . . . . . . Coherent radiation from a collection of Spectrum . . . . . . . . . . . . . . . . The general multipole expansion . . . Radiation Reaction . . . . . . . . . . . Example . . . . . . . . . . . . . . . . . Exercises . . . . . . . . . . . . . . . .

. . . . . partices . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

42 43 45 45 46 47 49

4 Radiation from Relativistic Systems 50 4.1 Four-vector Notation . . . . . . . . . . . . . . . . . . . . . . . . . 50 4.2 Specic Examples of Lorentz Transformations . . . . . . . . . . . 52 4.3 Electromagnetism . . . . . . . . . . . . . . . . . . . . . . . . . . . 54 4.4 Emission from Relativistic Particles The Relativistic Larmor formula . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57 4.5 Further Invariant quantities . . . . . . . . . . . . . . . . . . . . . 58 4.6 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59 5 Bremsstrahlung 5.1 Qualitative Discussion of Bremsstrahlung . . 5.2 The Straight Line Approximation . . . . . . . 5.3 Multiple encounters . . . . . . . . . . . . . . 5.4 Gaunt factors . . . . . . . . . . . . . . . . . . 5.4.1 Classical corrections . . . . . . . . . . 5.4.2 Quantum mechanical corrections . . . 5.4.3 Bremsstrahlung in a Thermal Plasma 5.5 Collisional Damping . . . . . . . . . . . . . . 5.6 Other Considerations . . . . . . . . . . . . . . 5.7 Questions . . . . . . . . . . . . . . . . . . . . 6 Synchrotron Emission Theory 6.1 Emission by Relativistic Particles . . . . . 6.2 Semiquantitative Treatment . . . . . . . . 6.3 The Emissivity for Synchrotron Emission 6.4 Emission by a Power-Law Distribution . . 6.5 Synchrotron Self-Absorption . . . . . . . . 6.6 The Razin Eect . . . . . . . . . . . . . . 6.7 Cyclotron emission . . . . . . . . . . . . . 6.8 Questions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60 60 61 63 64 64 65 65 66 67 67 70 70 73 75 77 78 80 81 82 84 84 85 88 88 89 89

7 Astrophysical Synchrotron Sources 7.1 Summary . . . . . . . . . . . . . . . . . . . 7.2 The spectrum due to synchrotron radiation 7.3 Simple source models . . . . . . . . . . . . . 7.3.1 Slab geometry . . . . . . . . . . . . 7.3.2 Core-jet model . . . . . . . . . . . . 7.4 Evolution of Synchrotron Spectra . . . . . .

7.5

7.4.1 Synchrotron lifetime in astrophysical sources . . . . . . . Questions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

93 93 95 95 97 99 101 101 102 103 106 107 108

8 Compton Scattering 8.1 Thomson Scattering . . . . . . . . . . . . . . . . . . . . . . 8.1.1 Compton Scattering . . . . . . . . . . . . . . . . . . 8.2 Inverse Compton Scattering . . . . . . . . . . . . . . . . . . 8.2.1 Average scattered power . . . . . . . . . . . . . . . . 8.3 Synchrotron Self-Comptonised Emission . . . . . . . . . . . 8.3.1 The Inverse Compton Limit for Synchrotron Sources 8.4 Eectiveness of Compton Scattering . . . . . . . . . . . . . 8.5 Repeated scatterings for < 1 . . . . . . . . . . . . . . . . 8.6 Kompaneets equation . . . . . . . . . . . . . . . . . . . . . 8.7 Questions . . . . . . . . . . . . . . . . . . . . . . . . . . . .

9 Propagation Eects in Astrophysical Plasmas 109 9.1 The plasma response tensor . . . . . . . . . . . . . . . . . . . . . 110 9.2 Dispersion smearing . . . . . . . . . . . . . . . . . . . . . . . . . 113 9.3 Faraday rotation . . . . . . . . . . . . . . . . . . . . . . . . . . . 114 9.4 Interstellar Scintillation . . . . . . . . . . . . . . . . . . . . . . . 115 9.4.1 Propagation through an inhomogenous medium: The Parabolic Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . 116 9.4.2 Intensity uctuations quantitative . . . . . . . . . . . . 122 9.4.3 Scintillation regimes . . . . . . . . . . . . . . . . . . . . . 124 9.4.4 Scintillation of non point-like (extended) sources . . . . . 128 9.5 Questions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130 10 Exercises 132

Radiation Astrophysics 2010


Timetable

Week
1 2 3 4 5 6

Date
29 March 12 April 19 April 26 April 3 May 10 May

Topics covered
Radiative transfer, thermal radiation, Einstein co-es. Maxwells equns, Polarization, EM potentials, radiation from moving charges Special relativity, Bremsstrahlung Synchrotron radiation, Synchrotron sources Compton scattering, EM wave propagation through plasmas Review

Jean-Pierre Macquart J.Macquart@curtin.edu.au Tel: 9266-9248

Physical Constants and Conversion between CGS and SI units


Astronomical texts and articles frequently express quantities in cgs (gaussian) units rather than SI units (i.e. length in cm rather than m, charge in esu rather than coloumbs, energy in erg rather than joules, etc.) It is important to be able to use both sets of units interchangeably. Here we provide a list of physical contants in both units and describe how to convert formulae between the two unit systems. Table 1: Physical Constants constant symbol SI units speed of light c 3.0 108 ms1 electron charge e 1.6 1019 C electron mass me 9.1 1031 kg proton mass mp 1.67 1027 kg Boltzmann constant B 1.38 1023 JK1 electron volt eV 1.6 1019 J gravitational constant G 6.67 1011 Nm2 kg2 Planks constant/2 h 1.05 1034 J s classical electron radius re 2.8 1015 m Thomson cross section T 6.65 1029 m2 critical B eld Bc 1.44 109 T permittivity of free space 0 8.85 1012 F m1 permeability of free space 0 1.23 106 H m1

gaussian units 3.0 1010 cms1 4.8 108 esu 9.1 1028 g 1.67 1024 g 1.38 1016 JK1 1.6 1012 erg 6.67 108 dyne cm2 g2 1.05 1027 erg s 2.8 1013 cm 6.65 1025 m2 1.44 1013 G

Table 2: Astronomical Constants quantity symbol SI units gaussian units astronomical unit AU 1.5 1011 m 1.5 1013 cm parsec pc 3.09 1016 m 3.09 1018 cm light year ly 0.95 1016 m 0.95 1018 cm angstrom 1010 m 108 cm A solar radius R 7.0 108 m 7.0 1010 cm Earth radius RE 6.4 106 m 6.4 108 cm Jovian radius RJ 7.1 107 m 7.1 109 cm solar mass M 2.0 1030 kg 2.0 1033 g solar luminosity L 3.9 1026 W 3.9 1033 erg s1

Table 3: Conversion Factors between SI and gaussian units quantity length mass energy power force charge electric eld current current density magnetic induction SI/gaussian units 102 m/cm 103 kg/g 107 J/erg 107 W/erg s1 105 N/dyne 1 9 statcoul/C 3 10 4 3 10 V m1 /statvolt cm1 1 9 A/statamp 3 10 1 5 10 A m2 /statamp cm2 3 4 10 T/G gaussian/SI 102 cm/m 103 g/kg 107 erg/J 107 erg s1 /J 105 dyne/N 3 109 C/statcoul 1 4 statvolt cm1 /V m1 3 10 9 3 10 statamp/A 3 105 statamp cm2 /A m2 104 G/T

The formulae in the course lecture notes are given in SI units, but the formulae and quantities appearing in Rybicki & Lightman are expressed in gaussian units. The conversion of a formula to gaussian units involves no change when no electromagnetic quantities are involved. Furthermore, the following electromagnetic quantities do not change:

charge, q charge density,

electric eld, E electric potential, .

current density, J

The following quantities do need to be changed (the direction of the arrow is for conversion from SI to gaussian units)

magnetic eld B B/c permittivity of free space 0 1/4 vector potential A A/c permeability of free space 0 4/c2 .
For example, the formula F= is equivalent to F= q1 q2 r r3 (gaussian units). 1 q1 q2 r 40 r3 (SI units)

In cgs (cgs=centimetres, grams, seconds), the quantities E , B , D and H all have the same units (see qu. 7 below). The permittivity and permeability of free space in cgs units are equal to unity: 0 = 0 = 1. The unit of charge in cgs units is the statcoloumb: 1 coloumb = 3 109 statcoloumb (2) (1)

Exercises
Convert the following quantities and formulae from SI units to gaussian units 1. v = 10 m s1 2. B = 0.1 T 3. Plasma frequency, p = 56.4ne
1/2 1

4. Alfv en speed, vA = 2.2 1016 B ( ne )1/2 m s1 5. Ion sound speed vs = 91Te


1/2

m s 1
1/2

6. Thermal electron speed Ve = 3.9 103 Te 7. Maxwells equations D = B E H where D = B = = = =

m s 1

0 0 B t 0 J + 1 D , c2 t

(3) (4) (5) (6)

E H.

(7) (8)

Verify that E , B , D and H all have the same units in the cgs units and derive the relation between the cgs units commonly used to measure magnetic eld and those units commonly used to measure electric eld strength.

Chapter 1

Fundamentals of Radiative Transfer


Reference: Rybicki & Lightman, Chapter 1
In this chapter we are mainly concerned with the transfer of radiation in the geometric optics limit, which is to say that we ignore its wave properties. We regard the propagation of electromagnetic radaition as being associated with photons which propagate like rays. We will be more specic about the restrictions under which the radiative transfer approach is valid at the end of the chapter.

1.1

The Electromagnetic Spectrum

The electromagnetic spectrum covers a large range in energy, from < 106 eV at radio wavelengths to > 106 eV at -ray wavelengths. The are a number of equivalent ways to represent the same radiation: in terms of its wavelength (or wavenumber k = 2/), frequency (or angular frequency = 2 ), energy or temperature. In a vacuum the product of the wavelength and frequency equal the speed of light k = = c c (1.1) (1.2)

The energy of each photon of frequency is E = h = h , (1.3)

and its temperature can be dened by equating it to an energy using Boltzmanns constant E = kT. 9 (1.4)

1.2

Geometric considerations

In this course we will often consider radiation interesecting a sphere centred on a radiating source. The solid angle subtended by a section of this sphere is d = sin dd sterradians (abbreviated ster).

whole sphere has 4 ster of solid angle

dS=sinR2d=R2 sindd R2 d2

Figure 1.1: Solid angles over the surface of a sphere. For small angle ranges the surface corresponding to the solid angle is nearly at. For example, the surface area subtended by a small region of angle d on a sphere of radius R is (Rd)2 . However, as d becomes large, we do NOT get the right answer (e.g. see what happens if we set d = 2 ). The correct element of surface area is dS = R2 d.

1.2.1

Specic intensity

Consider a hypothetical telescope collecting radiation from a source. We dene the energy ux, F as the amount of radiation crossing perpendicular to the telescope surface per unit time per unit area. The energy ux is a vector which : points in the direction perpendicular to the surface, n F = Fn J s1 m2 . (1.5)

Thus, the total energy crossing an element of area A on the telescope surface is dE = F dA dt (1.6)

The ux depends on the distance to the emitting source. In the absence of any intervening absorption or emission, ux is a conserved quantity. For an isotropic source, the area-integrated uxes crossing spheres of 10

radii R1 and R2 centered on the source are equal. dE1 = dt 2 4R1 F (R1 ) = dE2 dt 2 4R2 F (R2 ). (1.7) (1.8)

So for an isotropic source the ux follows an inverse square law F (R) = F (R0 )
2 R0 constant = . 2 R R2

(1.9)

Figure 1.2: Some basic relations. We can dene a quantity that is intrinsic to the source and takes into account the possibility that the source does not emit isotropically or that it radiates dierent amounts of energy at dierent frequencies. We call this specic intensity or specic brightness, I , which is the amount energy crossing a surface of unit area per unit time per solid angle per frequency. Again I points in the direction of perpendicular to the collecting surface I = I n J s1 m2 ster1 Hz1 . (1.10)

The amount of energy crossing per a surface of dA over a time interval dt over solid angles d with observing bandwidth d is dE = I dA dt d d. (1.11)

The specic ux, F , is the specic intensity integrated over solid angle F = I d. (1.12)

11

This is the ux carried by photons through a solid angle d for a range of frequencies between and + d over a surface whose normal points in the . direction n In general the specic intensity depends on the frequency and direction I = I (, , ). The specic intensity of a ray travelling (in a straight line) in a vacuum is constant, and does not vary along the ray path: dI = 0. ds (1.13)

Consider the radiation from an isotropically emitting source. Each ray has specic intensity I . Suppose the source is spherical with radius r, so that the angle subtended by the source a distance R away is c = sin1 r/R.

Figure 1.3: What is the ux a distance R from the sphere? Using (1.12), the ux received at a point P a distance R away from the source is F = d I n = = = = I I
0

I cos d cos (sin dd)


2 c

d
0

d cos sin . (1.14)

r R

So we recover the inverse square law for the decrease in specic ux with distance from the source. But while the specic ux decreases with distance, the specic intensity does not.

1.2.2

Net ux, momentum ux and radiation pressure

We can equate the energy associated with the photons to a momentum and pressure. A familiar manifestation of this in the physical world is in the tails of comets. Comets commonly have two tails, one of which points in the direction of the solar wind (a stream of particles given o by the sun) and the other whose tail points directly away from the sun. This latter tail is due to the radiation pressure exerted by the suns radiation on the particles given o by the comet. 12

From Einsteins theory of special relativity, the momentum carried by a 4 particle of energy E with rest mass m0 is E 2 = p2 c2 + m2 0 c . The momentum direction is carried by a (massless) photon travelling in the z p= E z. c (1.15)

We can equate this to a force and pressure, P = dF/dA using Newtons law F = ma = dp dt . The momentum ux dp/(dtdA) = dE/(c dtdA) is exerted parallel ), and we can dene a vectorial to the direction of the rays (i.e. parallel to z pressure element, dp dP z= dF I cos d z dp dF = , z z = = dA dtdA c c (1.16)

z = cos is the angle between the direction of ray propagation and where n the normal to the surface which the photons are crossing ( n). Note that dp is a vector. In calculating the component of the momentum exerted perpendicular to the surface, we need to take the component of the momentum ux that is normal to the surface, i.e. we need dp n. This is dp cos , and the total momentum ux per unit area on the surface is p = 1 c I cos2 d N m2 Hz1 . (1.17)

Note that the momentum ux per unit area is a pressure, so p has the dimensions of pressure per unit frequency: it is the pressure exerted on a surface due to photons in the frequency range to + d . Radiation pressure is analogous to gas pressure. It is a scalar for isotropic radiation elds; the radiation only exerts a force when there is a gradient in the photon pressure.

Figure 1.4: Radiation pressure in action.

13

Figure 1.5: Absorption by a uniform medium.

1.2.3

Radiative energy density

The specic energy density u (units J m3 Hz1 ) is the energy per unit volume of space per unit frequency. Consider how much energy is contained a cylinder of volume dV = dA c dt of area dA at its ends and with length dx = c dt due to rays propagating over a range of solid angles d parallel to the length of the cylinder dE = I dA dt d d = I = u () dV d d, dV d d c (1.18)

where the second line follows from the denition of specic energy density. Thus the specic energy density for the rays propagating over a range of solid angles d about the ray parallel to the length of the cylinder is u () = I . c (1.19)

Thus the total energy density in the cylinder per unit frequency is found by integrating over all possible ray directions: u = u d = 1 c I d. (1.20)

The total energy density is found by integrating u over all frequencies. The mean intensity, J is dened as the specic intensity averaged over all solid angles: J = so that u = 4 J . c (1.22) 1 4 I d W m2 s1 Hz1 , (1.21)

The mean intensity is a property of a volume of space, and does not depend on a direction. The numerical values of the mean intensity and specic intensity are equal if the radiative source emits isotropically (i.e. J = I ). 14

Example: radiation pressure from an isotropic photon eld


As an example, lets calculate the radiation pressure due to an isotropic radiation eld, such as the relic 3K microwave background radation from the Big Bang. Here we have J = I , and the radiation pressure (eq. (1.17)) is found by integrating over the sphere from = [0, ] and = [0, 2 ] and p = = 1 1 I cos2 d = J c c 1 4 J = u 3 3
2

d
0 0

d cos2 sin dd (1.23)

Compare this with the relation for an ideal gas, where p = 2/3u. Why the dierence?

1.3

Radiative Transfer

So far we have only considered the properties of radiation emitted from a source without any propagation eects. The most obvious propagation eects are absorption and emission and scattering. It is most useful to treat these processes by considering changes in the specic intensity, I , whose value remains otherwise constant along the ray path. Below we treat eects due to emission and absorption, but defer a discussion of scattering until later.

1.3.1

Emission

We can dene an emission coecient, j (units J s1 m3 ster1 Hz1 ), as the rate at which energy is emitted by a medium between frequencies and + d , per unit volume per solid angle ( to + d) per time: dE = j dV d dt d (1.24)

When a beam travels a distance ds through an emitting region it sweeps out a volume dsdA, so the change in energy between frequencies and + d is, after relating the energy to the specic intensity (eq. (1.11)) dI = j ds. (1.25)

The integral of j over all angles yields the total power radiated by a medium per unit volume per frequency. In the case of the isotropic emission, this total power is P = 4j . We can equivalently describe the emission in terms of an emissivity, (units W kg1 Hz1 ), which is the power radiated per unit mass of material between frequencies and + d and integrated over all angles. For isotropic emission one has dE = dV dt d 15 d . 4 (1.26)

The emission coecients and emissivity are thus related according to j = . 4

(1.27)

1.3.2

Absorption

Unlike emission, the amount of absorption occuring in a medium depends on the intensity of the incident radiation. There obviously cannot be any absorption if no energy is incident upon the absorbing medium. We dene the absorption coecient, , (units m1 ) as the rate at which the fraction of the total intensity (i.e. dI /I ) decreases per unit distance along the ray path at a given frequency: dI = I ds. (1.28)

Note the sign of the absorption coecient: a positive value of implies that the ray looses energy along the ray path. A negative absorption coecient corresponds to stimulated emission (i.e. maser action), and leads to an increase in the specic intensity. We can also dene an absorption cross-section, (units m2 ) of a single particle in the absorbing medium. For a particle density (particles per unit volume) in the absorbing medium the absorption cross section is related to the absorption coecient by = . (1.29)

The amount of energy absorbed by the medium from a beam of radiation with specic intensity I is E = I dV d dt d. = I dA ds d dt d (1.30) (1.31)

Equivalently, we can describe the absorption in terms of an opacity, = n u/ (units m2 kg1 ), which describes the loss of intensity in terms of the mass of absorbers encountered per unit length along the ray path. Caveats Several conditions must be satised in order for the foregoing physical descriptions to be valid: i The absorbing particles must be randomly spatially distributed (or, more correctly, homogeneously distributed on suciently macroscopic scales). ii The absorption of each particle must be independent of the absorption of all other particles. (i.e. the absorbed energy scales linearly with the number of particles) iii There must be no shadowing, meaning that the absorption cross section must be small compared to the interparticle distance: 1/2 d 1/3 . This is related to point (ii). 16

1.3.3

The equation of radiative transfer

Combining the eects due to absorption and emission we have the equation of radiative transfer dI = j I . ds (1.32)

Note that in deriving this equation we are implictly assuming that the radiation propagates like a bundle of rays. It thus neglects the wave nature of the electromagnetic radiation, and treats neither reection, refraction nor diraction. There are many astrophysical situations where this treatment is inappropriate, particularly at radio wavelengths. For instance, it cannot be used to treat scintillation-related eects (such as the twinkling of starlight) and the ducting of radio waves within dense environments like pulsar magnetospheres. Solutions of the Transfer Equation Let us explore the implications of the radiative transfer equation. In the absence of absorption, the specic intensity a distance s along the ray-path is
s

I (s) = I (s0 ) +
s0

j (s ) ds .

(1.33)

Thus, the intensity increases linearly with distance.


I

Figure 1.6: Intensity as a function of distance for emission only. In the absence of the emission the specic intensity is
s

I (s) = I (s0 ) exp

(s ) ds

(1.34)

s0

Note that, whereas the intensity increases linearly with distance for homogeneous emission, it decreases exponentially for homogeneous absorbing medium. It is convenient to dene the dimensionless optical depth, through an absorbing medium in terms of the absorption coecient an the length of the ray path:
s

(s) =
s0

(s ) ds , 17

(1.35)

Figure 1.7: Intensity as a function of distance for absorption only. so that the intensity in a purely absorptive medium is written I (s) = I (s0 )e (s) . (1.36)

Thus the intensity decays by a factor of e1 when the ray has moved a distance corresponding to an optical depth of = 1. For a typical photon of frequency propagating through an absorbing medium, most photons escape the absorbing medium when < 1, and the medium is called optically thin. However, when > 1, most photons are absorbed, and medium is called opaque or optically thick. General Solution We can construct a general solution to the equation of radiative transfer. To do this, it is convenient to dene a source function S (units W m2 ster1 Hz1 ): S (s) = j (s) . (s) (1.37)

When multiple processes contribute to the emission and absorption the source function is
tot S =

(1.38)

This is relevant, for example, when considering absorption on a spectral line, in which there is both continuum and spectral line absorption. We can now write the radiative transfer equation as dI = S I . d This has a general solution (see exercise 2.) I ( ) = I (0)e +
0 e( ) S ( ) d .

(1.39)

(1.40)

18

As an example, let us consider the variation in intensity with path length through a homogeneous medium (i.e. S = const) in which both absorption and emission occur. Then the specic intensity is I ( ) = I (0)e + S 1 e . I the medium is optically thin, this solution simplies to I ( ) I (0)[1 ] + S . thin (1.42) (1.41)

On the other hand, in limit of an optically thick medium, , the solution reduces to I ( ) S . (1.43)

What is the average optical depth seen by an individual photon as it propagates from = 0 to = ? To determine this, we integrate the optical depth over the ray path, weighted by the number of photons remaining at that optical depth, e , and normalised by the total number of photons: =
0

e d = 1. d 0 e

(1.44)

Thus the average photon travels through a medium a distance corresponding to an optical depth of 1 before being absorbed. This distance is called the mean free path, l : = l . (1.45)

In other words, the mean free path is the inverse of the absorption coecient: 1 1 l = . = ( )

1.4

Thermal radiation

Here we consider radiation from a body that is in local thermodynamic equilibrium (LTE) with its surroundings. For matter in LTE all atomic, ionic and molecular level populations are described entirely by the local temperature (and are given by Saha-Boltzmann statistics: Ni /N exp(Ei /kT )). If we immerse a container in a bath of temperature T and wait for it to reach equilibrium with its surroundings, and then drill a small hole through the container to sample its radiation, what will we see? The specic intensity, I , of this radiation will depend only on T , and not on the shape or size of the container or any other property. If we connect two containers at temperature T the specic intensity must be the same from each container (i.e. I 1 = I 2 ) otherwise this would imply nett energy ow between two bodies of equal temperature, thus violating the second law of thermodynamics. In the same way, the radiation

19

Figure 1.8: Two containters at the same temperature, T , connected by a thin pipe. must also be emitted isotropically. Thus, because I depends only on T (and ) for blackbody radiation we write B (T ) = I . (1.46)

Now let us now consider the two connected containers as one body and consider their blackbody emission as a system. Because I = B is constant everywhere (i.e. dI /d = 0), the equation of radiative transfer now implies dI = S (T ) B (T ) = 0. d j = B (T ), (1.47)

Thus, for a blackbody the source function S = B (T ), implying Kircho s Law (1.48)

which is to say that a blackbody is both a good emitter and absorber, and that the two are related. At this stage it is appropriate to point out that Kirchos law holds for all thermal radiation, but that not all thermal radiation is blackbody radiation. Thermal radiation only becomes blackbody radiation for optically thick media. When this is not the case, then I = B , and dI /ds = 0. However, since S = B (T ) still holds, we still have j = B (T ), and the transfer radiation for thermal radiation is dI = I + B (T ). (1.49) ds The right hand-hand side of this equation is zero only for blackbody radiation. Can you see why it might not hold for optically thin media?

1.4.1

Thermodynamics of blackbody radiation

Recall that the change of heat dQ in medium is related to its change in internal energy dU and the work done on the medium p dV according to the equation dQ = dU + pdV. 20 (1.50)

Also recall that the entropy change of the medium, dS , is related to its change in heat by dS = dQ . T (1.51)

Now the total internal energy of radiation is given by U = uV , where we derived p = u/3 for isotropic radiation. For blackbody radiation the internal energy is governed only by the temperature u = u(T ). Combining this with equations (1.50) and (1.51) we have dS = V du udV udV V du 4u dT + + = dT + dV. T dT T 3T T dT 3T (1.52)

Since the temperature and volume are independent quantities for blackbody radiation, equation (1.52) actually implies two independent equations S T We thus have 2S 1 du 2S 4 = = = T V T dT V T 3 Rearranging, we have dT du =4 , u T the solution of which is u(T ) = aT 4 , where a is a constant to be determined. Now since u = B (T ) = Similarly, the ux is F = T 4 , where a = = 4 = 7.56 1015 erg cm3 deg4 c ac = 5.67 105 erg cm2 deg4 s1 4 (1.59) (1.60) (1.61) (1.58) ac 4 T . 4
4 c

=
const V

V du T dT

and

S V

=
const T

4u . 3T

(1.53)

u 1 du + T 2 T dT

(1.54)

(1.55)

(1.56) B (T )d , we have (1.57)

21

[Exercise: convert these to SI units!] You should also be able to show from the foregoing equations that the following relations hold for radiation S T V 1/3 pV
4/3

= = =

4 3 aT V 3 constant constant.

(1.62) (1.63) (1.64)

and compare them to the corresponding relations with an ideal gas.

1.4.2

The Plank Spectrum

What is the exact shape of a blackbody spectrum? To derive this we need to employ some elementary quantum mechanics and to know (i) the number density of photon states for a given energy level and (ii) the average energy of each state. We consider the blackbody as a box of volume V lled with radiation of all wavelengths, with associated wavevectors k= 2 . n (1.65)

The photon number density for each element of d3 k is thus NS = 2 1 1 2 d d3 k 2 = 2 k dk d = 2 d, (2 )3 (2 )3 (2 )3 c3 (1.66)

where the extra factor of 2 appears because a photon may have one of two possible polarization states for each specic k. The density of states per unit solid angle between and + d is S = 2 2 . c3 (1.67)

The average energy of photons in each state is determined using the fact that photons are indistinguishable from one another, and thus obey Bose-Einstein statistics. Thus the occupation number of energy level En = nh is proportional to exp(En /kT ). So the average energy of photons with frequency is just a sum over all possible energy levels E =
En /kT 0 En e En /kT 0 e

h . exp(h/kT ) 1

(1.68)

The energy density at frequency , u , is a product of the number of photons with energy E with their density, and since u c = B , we have B (T ) = B (T ) = 2h 3 /c2 exp(h/kT ) 1 2hc2 /5 . exp(hc/kT ) 1 22 (1.69) (1.70)

The hotter the blackbody, the bluer the peak of the radiation (see Ex. 3) and the higher the value of B at any given : blackbody curves at dierent T do not cross. As an exercise derive the Stefan-Boltzmann equation directly from the Plank spectrum: F = F d = B dT 4 , (1.71)

where the Stefan-Boltzmann constant, is given by = Rayleigh-Jeans limit: h kT At suciently low frequencies the energy of each photon is much smaller than the average kinetic energy of the particles radiating inside the blackbody. In this case we can expand the Plank spectrum for h/kT 1, which gives us the specic intensity in the Rayleigh-Jeans part of the spectrum I (T ) = 2 2 kT. c2 (1.73)
4 2 5 kB . 2 15c h3

(1.72)

The properties of radiation in this regime are well described by classical physics. We could have derived the specic intensity using I = S c E classical with the average energy of each particle E classical = kT . Note that this approximation fails once quantum mechanical eects start to become important (i.e. when the energy of the photon being considered is comparable to the energy of the individual radiating particles). We can see that the form of I specied in the Rayleigh-Jeans approximation blows up as . This is called the ultra-violet catastrophe. Wien limit: h kT When the energy of the photon being considered exceeds the typical kinetic energy of the radiating particles the Plank spectrum takes the form I (T ) = 2h 3 h exp 2 c kT . (1.74)

The specic intensity falls rapidly in the Wien part of the spectrum.

1.4.3

Temperatures associated with emission

The Brightness Temperature Not all astrophysical objects radiate like blackbodies, however this does not stop people assigning brightness temperatures to the emission. This is done by 23

choosing a temperature T , such that the observed specic intensity, I , matches the specic intensity that of blackbody radiation at that frequency: T such that I = B (T ). (1.75)

However, unless the object actually is a blackbody, the exact value of T will vary with frequency. For example, synchrotron emission is a distinctly non-thermal emission process, but it is often useful to equate emission observed from this process with a brightness temperature. Here, in the Rayleigh-Jeans limit we have Tb = c2 F c2 I = . 2 2 k 2 2 k (1.76)

b Now since dI/d = I + B (T ) it follows that dT d = Tb + T , so the brightness temperature depends on the optical depth according to

Tb ( ) = Tb (0)e + T 1 e , Eective temperature

h kT .

(1.77)

We can also equate a temperature with the total ux received from a source:
4 Te such that F = Te .

(1.78)

The eective temperature is the temperature for which the total received ux satises the Stefan-Boltzmann law.

1.5

Einstein coecients

So far we have only described the properties of the radiation in terms of macroscopic physical quantities. What do emission and absorption correspond to on a microscopic scale? Consider transitions between two energy states separated by an energy E = h0 .

Figure 1.9: A two level atom.

24

We can dene a probability associated with transitions between levels. Emission corresponds to a transition from level 2 down to level 1. Let us dene A21 = probability for spontaneous emission per unit time. Similarly, a transition from level 1 to level 2 corresponds to absorption, and requires external photons to supply the energy to make the transition. We write = B12 J probability for absorption per unit time due photons with energy E h0 , = where J
0

J ( ) d,

(1.79)

and where ( ) is the shape of the line prole and J is the mean intensity of the photon eld at frequency .

Figure 1.10: The line prole ( ). The system can also move from level 2 to level 1 via stimulated emission, which is a process that requires external photons whose energy corresponds to the energy dierence between levels. You can think of these external photons tickling at the resonant frequency of the atom, forcing it to emit its own photon of the same energy. Stimulated emission is the principle behind the operation of a laser. We can dene a probability associated with this negative absorption = transition probability per unit time associated with stimulated emission. B21 J (1.80) Now, if the system is thermodynamic equilibrium, there are just as many transitions from level 1 to 2 as there are from level 2 back down to 1. So if the number density of atoms in states 1 and 2 are n1 and n2 we have = n2 A21 + n2 B21 J, n1 B12 J = J
A21 B21 n1 n2 B12 B21

(1.81) (1.82)

. 1

25

But we have an additional piece of information about the density of states when the system is in thermodynamic equilibrium. In particular, n1 g1 exp(E/kT ) = . n2 g2 exp((E + h0 )/kT ) (1.83)

where g1 and g2 are statistical weights associated with each level. Substituting this in, we have, for thermal radiation = J
A21 B21 g1 B12 g2 B21

exp(h0 /kT ) 1

= B .

(1.84)

But we saw above that B has a Plank spectrum. By equating B with the Einstein coecients in the above equation, we have related the emission and absorption to microphysical properties of the radiating atoms. This yields the Einstein relations g1 B12 A21 = = g2 B21 2h 3 B21 . c2 (1.85) (1.86)

So we see that the absorption and emission probabilities are related. These probabilities are constants for any particular atom, and do not depend on the temperature. We can relate the Einstein coecients back to their macrophysical emission and absorption counterparts. We assume that the system both emits and absorbs with the same line prole ( ). For emission at frequency , we have dE = j dV dddt = j = h ( )n2 A21 dV dddt. 4 (1.87) (1.88)

h n2 A21 ( ). 4

The corresponding equation for absorption is dE = I dV dddt = a = h0 (n1 B12 n2 B21 ) I dV dddt (1.89) 4 (1.90)

h ( )(n1 B12 n2 B21 ). 4

The nal term, n2 B21 , is the correction for stimulated emission. We can combine these expressions with the Einstein relations to show that (exercise!) S = 2h 3 j = 2 c g 2 n1 1 g 1 n2
1

(1.91)

26

1.5.1

Maser emission

Now, if the matter is in thermal equilibrium with itself (local thermodynamic equilibrium), though not necessarily with the surrounding radiation, we can use eq. (1.83) to show that the absorption coecient is = h n1 B12 [1 exp(h/kT )] ( ), 4 n2 g 1 h ( ), n1 B12 1 4 n1 g 2 n2 n1 < . g1 g2 (1.92)

which is always positive. More generally, however one has = (1.93)

which is negative when the population is inverted, (1.94)

In this case, stimulated emission dominates over normal absorption, and the intensity of the emitted radiation increases proportional to the intensity of the incident radiation. The emitted radiation is emitted in the same direction and with the same phase as the incident radiation. This is known as MASER emission (Microwave Amplication by Stimulated Emission of Radiation).

1.5.2

Line proles

In the preceeding section we talked about the line prole ( ) without specifying how this is determined. Na vely, you might think that the line prole is just a Dirac delta function centered on the frequency corresponding to the energy of the transition. However, both quantum mechanical and classical considerations ensure that this is not the case. A more careful treatment of the physics associated with this awaits in Astrofysica B, but for the time being we merely mention the processes that contribute to line broadening. Recall from the Heisenberg uncertainty principle that the uncertainty in the energy of an energy level E is related to the uncertainty in the time spent in that state by E h . t (1.95)

Thus the uncertainty in the energy dierence E2 E1 is approximately the sum of the uncertainties associated with each of the individual energy levels E21 = E1 + E2 = h h + , t1 t2 (1.96)

where t1,2 are the lifetimes in the initial and nal states. Thus, the width of the spectral line is approximately 2 2c 1 1 + t1 t2 27 . (1.97)

A more careful calculation shows that the line prole follows a Lorentzian form ( ) = /4 2 , ( 0 )2 + (/4 )2 (1.98)

where = n n Ann is the spontaneous decay rate for level n, and is calculated by summing over Einstein emission coecients. Line broadening due to the Heisenberg uncertainty is an intrinsic property of any spectral transition, and is often called natural broadening. Doppler line broadening is due to the motions of the particles that are absorbing (or emitting) radiation. If the particles are in thermal equilibrium their speeds are described by a Boltzmann distribution, and the typical speed of a particle at temperature T is v = (2kT /m)1/2 . Thus the typical linewidth is given by / = v /c. A full integration over all possible velocities in the Boltzmann distribution gives a line halfwidth = 2 c 2kT 2 + vbulk ln 2, m (1.99)

where the term vbulk describes any bulk motion of the emitting material towards or away from the observer. Pressure broadening is also important in certain circumstances when collisions between atoms are frequent. These collisions potentially disturb the electron orbital structure of the radiation atoms, thus aecting transition energies. The amount of pressure or collisional broadening depends on the mean time between collisions, which is the time required for the particles, travelling at speed v , to travel one mean free path t0 lmfp . v (1.100)

The linewidth associated with collisional broadening is 2 1 . c t0 (1.101)

The pressure broadened width of a spectral line generated in a stellar atmosphere can be used to estimate the local density and thus the gravity of the star.

1.6

Exercises

1. Show explicitly that an integral of the surface area element dS = R2 d yields dS = 4R2 . (1.102)

sphere radius R 28

2. Show that the general solution of the radiative transfer equation given by dI = S I . d is I ( ) = I (0)e +
0 e( ) S ( ) d .

(1.103)

(1.104)

[Hint: as a rst step, multiply both sides of (1.103) by d e , then con struct propagation equations for I = I e and S = S e .] 3. Show that the peak frequency, max , of the blackbody spectrum B (T ) occurs when B (T ) Hence show that hmax 2.82kT. (1.106) = 0.
=max

(1.105)

This is known as the Wien displacement law, and it shows that the peak of the blackbody spectrum shifts linearly to higher frequency with temperature. 4. Verify the form of the Plank spectrum by lling in the intermediate steps in equation (1.68).

29

Chapter 2

Review of Electromagnetic Theory


Reference: Rybicki & Lightman, Chapter 2 Optical Coherence & Quantum Optics, Mandel & Wolf, 6.3

Most of the formalism presented in the previous lectures was based on purely phenomenological models of radiation. Here we develop a description of the radiation in terms of electromagnetic theory based largely on Chapter 1 of Melrose & McPhedran and Chapter 2 of Rybicki & Lightman. We use this to compute the power and polarization of radiation emitted by an accelerated charge, and to describe the propagation of radiation through media (see also Chapter 9 of these notes).

2.1

Maxwells Equations and the Poynting Vector

The electric eld E is dened as the force per unit charge acting on a test charge. If q is an arbitrarily small test charge so that the eld it generates is negligibly small, then the electric force acting on it is q E. The magnetic induction B cannot be dened in the same way as the electric eld because there is no magnetic counterpart of a charge. Instead B is dened in terms of the force acting on a current I. The force per unit length is I B. The source terms for the electric eld and magnetic induction are the charge density and the current density J respectively. The charge density is dened by considering an arbitrarily small volume about the point x at time t: (x, t) = lim q (inside the volume (x V/2, x + V/2)).
V0

(2.1)

The charge density is dened by considering an arbitrarily small circular area centred on the point x at time t and oriented such that that normal to the 30

surface of the circle is along the direction of J. The magnitude of J is equal to the charge per unit area cross the surface per unit time. The units of current density are A m2 . The elds E and B are related to each other and to and J by Maxwells equations: E = B B = B t 1 E c2 t Faradays induction law Amp` eres Law (2.2) (2.3) (2.4) (2.5)

0 J + 0 0

E = =

Gauss Law No magnetic monopoles.

The parameters 0 = 8.854 1012 F m1 and 0 = 4 107 H m1 are the permittivity and permeability of free space respectively. The farad (1 F = 1C V1 ) is the unit of capacitance and the henry (1 H = V s A1 ) is the unit of inductance. Maxwells equations listed above are completely general and apply when a medium is present or in free space. In the presence of a medium the charge and current densities are separated into induced and extraneous parts by writing = ind + ext and J = Jind + Jext , and the induced parts are identied as the response of the medium. It is possible to incorporate the responses into two new elds, the induction D and the magnetic eld strength H. Then (2.3) and (2.4) are replaced by H = D = Jext + ext . D t (2.6) (2.7)

In a vacuum D and H are related to E and B by D = 0 E and H = B/0 . In the presence of an isotropic medium the induced electric eld and magnetic eld are in the same direction as the external electric eld and magnetic induction, and one has D = E and H = B/.

2.1.1

Continuity equations

Maxwells equations imply three continuity equations. A continuity equation expresses the fact that some quantity is conserved. The continuity equations for some quantity Q has the general form W + F = 0, t (2.8)

where W is the density of Q (i.e. Q = d3 xW ) and F is the ux of Q. Equation (2.8) may be integrated over volume to yield an alternate form which expresses the rate of change in Q with the surface integral of the ux: Q t
S

F = 0. d2 S n 31

(2.9)

This may be interpreted as stating that the time rate of change of Q is balanced by the ux of Q crossing the surface into the volume. Any source or sink of Q would appear on the right hand side, specifying the rate per unit time per unit volume at which the quantity Q is generated inside the volume. The divergence of equation (2.3) and time derivative of (2.4) imply the charge continuity equation + J = 0, t (2.10)

which is to say that charge can neither be created nor destroyed. An equation expressing the continuity of energy is derived by evaluating (E B) and rewriting the result using Maxwells equations. This yields t 0 |E|2 |B|2 + 2 20 + E B 0 = J E. (2.11)

The term on the right hand side is interpreted as the mechanical work done by the electric eld. Recall that the equation of motion of a charged particle is dp/dt = q (E + v B). The rate of increase in energy is thus dE/dt = v dp/dt = q v E, which is the rate at which the electric eld does work on the particle. If there are n particles per unit volume then the work done by the eld is ndE/dt = J E. Hence the work done on the eld by the particles is J E. The 0 E 2 /2 and B 2 /20 terms in (2.11) represent the energies stored in the electric and magnetic elds respectively. The quantity E B/0 is the Poynting vector, which represents the electromagnetic energy ux. The third continuity equation implied by Maxwells equations involves electromagnetic momentum. However, we not need concern ourselves with this equation for the present course.

2.2

Electromagnetic Potentials

As a rst step in solving Maxwells equations it is often useful to introduce the scalar potential and the vector potential A as alternate descriptions of the electromagnetic eld in place of E and B. It is worthwhile to remember that an arbitrary electromagnetic eld may be described in a variety of mathematically equivalent ways, and the only feature special about the description in terms of E and B is their straightforward physical interpretation. The mathematical motivation for introducing the potential is to satisfy two of Maxwells equations identically. Equations (2.2) and (2.5) are replaced by E B = = A. A t (2.12) (2.13)

Thus we have reduced two of Maxwells equations to subsiduary equations, but we still have to solve two coupled partial dierential equations. We can 32

reduce this further by forming another subsiduary equation by taking the time derivative of (2.4) and using the charge continuity equation to remove from the equations. We then have three subsiduary equations and one main equation to solve which is, after some simplication is 1 2 1 2 A + 2 + ( A) = 0 J, c2 t2 c t or equivalently 2 + A= . t 0 (2.15) (2.14)

2.2.1

Gauge conditions

Relations (2.12) and (2.13) do not dene A and uniquely. Specically, for an arbitrary dierentiable function , any potentials A and of the form A

= +

t = A ,

(2.16) (2.17)

also satisfy (2.12) and (2.13). The freedom to make gauge transformations allows one to impose a gauge condition on the choice of potentials. Specic gauge conditions correspond to particular choices of the arbitrary function . Three specic choices are A = 1 +A = c2 t = 0 0 0 Coloumb gauge Lorentz gauge temporal gauge. (2.18) (2.19) (2.20)

For the time-varying elds we consider in this course it will prove most useful to use the Lorentz gauge. In this case (2.16) and (2.17) reduce to 1 2 2 = c2 t2 1 2 2 A = c2 t2 0 0 J. (2.21) (2.22)

Thus in the Lorentz gauge is generated by and A is generated by J, with the sources related to their potentials by dAlemberts equation. However, there is no clear separation in this gauge between electric and magnetic eects. One can show that dAlemberts equation leads to the following solutions for the potentials: (t, x) A(t, x) = = 1 (t , x ) d3 x 40 |x x | 0 J(t , x ) , d3 x 4 |x x | 33 (2.23) (2.24)

where t = t |x x | c (2.25)

is the retarted time. The retarded time is interpreted as the time t at which the eld is observed minus the light propagation time from the point x where the source was at time t. This is due to the fact that the electromagnetic eld propagates from the source to the point of observation at the speed of light. In other words, the eld measured by an observer at x is determined by the conguration of charges and currents not as they currently are, but as they were a time t ago equal to the light travel time between the source and observer.

2.3

Plane Electromagnetic Waves

Consider the solution of Maxwells equations in vacuo. With both and J equal to zero, the curl of (2.2) combined with (2.3) yields a propagation equation for the electric eld. The vector identity ( a) = ( a) 2 a allows us to manipulate this into the form of the wave equation: 2 1 2 c2 t2 E = 0. (2.26)

An identical equation holds for B. One can construct a general solution of the wave equation from a superposition of plane waves of the form E B = E0 exp[i(k r t)] = B0 exp[i(k r t)], (2.27) (2.28)

/ is the wavevector that where E0 and B0 are complex amplitudes, k = 2 n and = 2 is the angular frequency points along the direction of propagation n of the radiation. One obtains a set of relations governing the eld amplitudes by substituting back into Maxwells equations: i k E0 i k B0 = = = = 0 0 i B0 i 2 E0 . c (2.29) (2.30) (2.31) (2.32)

i k E0 i k B0

Equations (2.29) and (2.30) show that the electric and magnetic elds both oscillate transverse to the direction of propagation. This is true for plane waves propagating in vacuo but the elds may contain some component parallel to the direction of propagation when a medium is present. The remaining two relations then imply that E0 , B0 and k form a right-handed triad of mutually

34

perpendicular vectors, and imply the following equations E0


2

= =

cB0 c k .
2 2

(2.33) (2.34)

The rst states that, when considering propagation of waves, we need only consider the electric eld, since the magnetic eld is always orthogonal to it with a magnitude set by the magnetude of the electric eld. The second states that the phase velocity of the wave is equal to the speed of light in vacuo: vphase = c. k (2.35)

Note that this does not hold when a medium is present. For example, in an isotropic medium with refractive index n the phase speed is /k = n c which may exceed the speed of light. For instance, the refractive index of a plasma is 2 n2 1 p / 2 , which is less than unity. Note, however, that the group velocity vg = k (2.36)

refers to the speed through wave electromagnetic energy propagates through the medium, and does not exceed the speed of light. We can take the time average of the electric and magnetic eld solutions given by (2.27) and (2.28) to determine the magnitude of the Poynting vector: S = E B/0 = 0 c c 1 |B0 |2 = E0 B0 /0 = |E0 |2 , 2 20 2 (2.37)

where we have used the fact that the magnitudes of the electric and magnetic eld amplitudes are related. Similarly the time averaged energy density per unit volume is U = 0 |E0 |2 0 c2 |E0 |2 |B|2 |B0 |2 |B0 |2 0 |E|2 = = . + + = 2 20 4 40 2 20 (2.38)

The rate of ow of energy (i.e. the group velocity) is the mean Poynting vector normalised by the energy density, vg = which is equal to c in the vacuo. S U (2.39)

2.4

Polarization: The Mutal Coherence Matrix and Stokes Parameters

Consider a well-collimated, uniform, quasi-monochromatic beam of light with mean angular frequency . The monochromatic plane waves we considered 35

above is a good example of such a beam of light. Let us separate each complex amplitude in the vector E0 into an amplitude and phase: E0 (t) = (E0 x (t), E0 y (t)) = (|E0 x (t)|eix (t) , |E0 y (t)|eiy (t)) , (2.40)

where we have allowed for the fact that the amplitude E0 may be a slowly varying function of time if the signal is quasi-monochromatic (it is a constant if the signal is truly monochromatic). Then the electric eld associated with the beam of light varies as
t] Ex (t) = |E0 x |ei[x (t)+kr t] Ey (t) = |E0 y |ei[y (t)+kr .

(2.41)

In general E0 x and E0 y will be correlated to some extent. We can dene a matrix of correlations between the electric eld oscillations in the x and y planes: J= E0 x (t) E0 x (t) E0 ( t ) E 0 x (t) y E0 x (t) E0 y (t) E0 y (t) E0 y (t) (2.42)

Note that the trace of the matrix J, known as the coherence matrix of the beam is proportional to energy density of the radiation: trJ = Jxx + Jyy = |E0 |2 . (2.43)

The coherence matrix contains four independent quantities, two being the amplitude of the electric eld along the x and y directions and the other two being the amplitude and phase of the correlation between the x and y (complex) electric eld amplitudes. Note that the coherence matrix can be used to describe unpolarized radiation. When the radiation is unpolarized the phase functions x (t) and y (t) vary randomly as functions of time, so the time-average quantities Jxy and Jyx average to zero in this instance. The diagonal components of the coherence matrix are still non-zero in this case because the each component of electric eld is still correlated with itself. In astronomy an alternate but equivalent system of describing the polarization state of a beam is often employed. These are the Stokes parameters, I, Q, U and V . I refers to the total intensity of the radiation, Q and U measure the linear polarization along the two mutually orthogonal axes relative to the direction of propagation, and V measures the circular polarization. In terms of the notation used above, the Stokes parameters are dened as follows: I Q U V where (t) = y (t) x (t). 36 (2.48) = = = = Jxx + Jyy = |E0 x |2 + |E0 y |2
2

(2.44) (2.45) (2.46) (2.47)

i(Jyx Jxy ) = 2 |E0 x ||E0 y | sin (t) ,

Jxx Jyy = |E0 x | |E0 y | Jxy + Jyx = 2 |E0 x ||E0 y | cos (t)

Note that Q, U and V are all real, but may be negative. The total intensity is always real and positive. The radiation is polarized if any of Q, U or V are non-zero. The degree of polarization, P is dened such that P 2 I 2 = Q2 + U 2 + V 2 . (2.49)

2.5

Fourier relations for electromagnetic waves

How is the energy of the radiation related to its electric eld. For monochromatic radiation one obviously has E = h . But for the radiation to be truly monochromatic it must vary sinusiodally for an innite amount of time. In practice, the radiation is emitted (and received) over a nite time, and the duration of the pulse t is related to the resolution (or uncertainty) with which we can determine its spectrum 1 . t (2.50)

The shorter the pulse, the broader the range of spectral components it contains. To make things explicit, the Fourier transform of the electric eld is ( ) = E E (t) =

1 2

E (t)eit dt

(2.51) (2.52)

( )eit d. E

Parsevals theorem relates the time-integrated energy with spectral-integrated energy according to

E 2 (t)dt = 2

|E ( )|2 d.

(2.53)

Now the energy per unit time per unit area is given by the Poynting vector, S (t). Integrating over time, the energy per unit area associated with the radiation is thus dW = dA

S (t)dt = 20 c
0 0

( )|2 d. |E

(2.54)

Thus the energy per unit area per unit frequency associated with a pulse of radiation is dW ( )|2 . = 2c0 |E dAd (2.55)

One cannot talk about an instantaneous spectrum of the radiation because of the uncertainty in measuring the frequency of a signal over a nite duration embodied in (2.50). One loses all frequency information as the time interval decreases to 0 (i.e. instantaneous). However, if we measure the signal over a 37

suciently long time interval, T , we can reliably measure the components of its power spectrum on frequencies much greater than 1/T : dW 1 = 2c0 lim |E ( )|2 . T T dAddt (2.56)

2.6

The Geometrical Optics Limit

The ray or geometrical optics treatment that we relied heavily on in Chapter 1 is only applicable in certain situations. Rays are lines tangent to the direction of propagation of the electromagnetic energy. Loosely speaking, we can use the geometric optics approximation whenever the intensity of the radiation does not vary greatly over one wavelength. Geometric optics is therefore an excellent approximation for treating the propagation of gamma rays, but it often fails disastrously for treating the propagation of, say, radio waves. As a practical instance, consider the focussing of a beam of light by a lens. Geometric optics predicts that the intensity at the focal point of the lens is innite. This is clearly non-physical, and geometric optics fails in the focal region. In this case wave optics is important in a region of length around the focal point. The propagation of rays in the geometrical optics limit follow Hamiltons equations (k, x, t) dx = , dt k dk (k, x, t) = , dt x d (k, x, t) = . dt t (2.57)

In a stationary medium the third equation is not relevant.

2.7

Exercises

1. Derive equation (2.26) for both E and B directly from Maxwells equations. 2. What is the coherence matrix for circularly polarized radation? For radiation linearly polarized in the xdirection? For unpolarized radiation? 3. Show that the trace and determinant of the coherence matrix remain unchanged for any rotation of the x and y axes about the direction of propagation (along the z axis). 4. Express the elements of the coherence matrix J in terms of the Stokes parameters (I, Q, U, V ). 5. Prove using the properties of the coherence matrix that I 2 Q2 + U 2 + V 2 . (2.58)

38

6. Consider the propagation of rays in the geometrical optics limit using Hamiltons equations. In the paraxial approximation one separates x into z , r, with the z axis along k = k . Show that Hamiltons equations yield an expression for the angular deviation of the ray in terms of the refractive index variations n d(n ) = . dz r (2.59)

Hence show that the angular deviation of a ray if it encounters a phase gradient = d/dr orthogonal to the direction of propagation is = 1 . k (2.60)

39

Chapter 3

Electromagnetic Radiation from Moving Charges


Reference: Rybicki & Lightman, Chapter 3 Melrose & McPhedran, Chapter 19
Accelerated charges radiate electromagnetic energy and electromagnetic energy can interact with particels to accelerate them. In this chapter we develop the electromagnetic theory to calculate the energy and polarization of an accelerated charge.

3.1

Li enard-Wiechart Potentials

The goal of this section is to develop a formula which describes the power radiated by a non-relativistic particle in vacuo. We rst consider the potential due to a charge with arbitrary trajectory and use this to derive the Poynting vector associated with the electromagnetic eld of this particle. This enables us to write down an expression for the power radiated away in the electromagnetic eld.

3.1.1

The Lienard-Wierchert potentials


x = X(t). (3.1)

Suppose we have a charge q whose position at time t is described by

The particles trajectory varies as a function of time according to its initial position and velocity and any forces acting upon it. We assume that the charge is located on a point particle, so that the charge density is zero everywhere except at the point X(t), at which the density is innite. Since the total charge is q , we write the charge and current densities as (t, x) = q 3 (x X(t)), J(t, x) = q v(t) 3 (x X(t)). 40 (3.2)

Now the electric and vector potentials for an abritrary charge are (t, x) = 1 40 0 A(t, x) = 4 (t , x ) (t t |x x |/c) |x x | J(t , x ) (t t |x x |/c) . dt d3 x |x x | dt d3 x q 40 dt (t t |x X(t )|/c) |x X(t )| (3.3) (3.4)

For our moving charge this becomes (t, x) = A(t, x) = 0 q 4 dt (3.5) (3.6)

(t ) (t t |x X(t )|/c) X . |x X(t )|

Performing the integral over time leaves us with (t, x) = q 1 40 |x X(tr )| [x X(tr )] X (tr )/c (3.7)

A(t, x)

Potential from a charge at its past location at X(tr ) (tr ) X 0 q (3.8) 4 |x X(tr )| [x X(tr )] X (tr )/c Vector potential from current density at past time tr

where tr = t |x X(tr )|/c is the retarded time. Our equations are complicated by the fact that tr depends implicitly on the trajectory of the particle, described by X(tr ). Further simplication is possible, however, by choosing the co-ordinate axis so that the particle is at the origin at t = tr and writing |x| = r. Equations (3.7) and (3.8) then yield the Lienard-Wierchert potentials: (t, x) = q 40 (r x v/c) 0 q v A(t, x) = 4 (r x v/c) (3.9) (3.10)

As a simple example of these pontentials, consider the elds for v = 0. We then recover the Coloumb potential of a point charge, = q/40 r with A = 0.

3.1.2

The Poynting Vector

The electromagnetic elds associated with the moving charge are derived from the potentials: E = A/t, B = A. (3.11)

Evaluation of the elds is complicated by the fact that derivatives refer to a co-ordinate system whose origin depends on the retarded time, tr , which is an 41

implicit function of both position and time. We merely state the derivatives are evaluated using the results tr 1 = , t 1 x v/rc The electric eld is then E=q with B= xE E/c, =x rc (3.14) (x rv/c)(1 v 2 /c2 + x a/c2 ) (ar/c2 )(r x v/c) , 40 (r x v/c)3 (3.13) tr = x c(r x v/c) (3.12)

= x/r is the unit vector along the direction of x. The Poynting vector where x gives the ux in electromagnetic radiation FEM = EB x E E), = c0 (|E|2 x 0 (3.15)

The inductive parts of the eld give rise to a contribution to the Poynting vector that varies with r as 1/r4 , and the radiative parts give a contribution that varies as 1/r2 ; there are also cross terms that vary as 1/r3 . The inductive terms are associated with a ow of electromagnetic energy that remains localised near the particle. The radiative terms imply an irreversible ow of energy from the particle to innity. The radiative terms give FEMrad = q2 |(x rv/c)x a ar(r x v/c)|2 . x 16 2 0 c3 (r x v/c)6 (3.16)

3.1.3

Larmor Formula

The power radiated by a particle is calculated from the rate of energy ow to innity. We consider a large sphere centred on the moving origin. This is given by integrating the radial component of the Poynting vector over the surface of the sphere and letting the radius of the sphere tend to innity. For a non-relativistic particle the ux in electromagnetic radiation is FEMrad = q2 q2 2 2 ( . | x x a a r | |x x a)|2 x x = 16 2 0 r6 c3 16 2 0 r2 c3 (3.17)

The contributions from the electric and magnetic elds are Erad Brad = q [ x ( x a)] 40 rc2 Erad . = x c 42 (3.18) (3.19)

Note that the angular distribution of the power due to this dipole radiation is proportional to sin2 P () = q 2 a2 , sin2 x 16 2 0 r2 c3 (3.20)

where is the angle between the acceleration and the viewing direction (i.e. the wavevector). Electric dipole radiation is linearly polarized along the projection of the axis of vibration onto the transverse plane. The total power is given by integrating the magnitude of FEMrad over the surface of a sphere. This is easily achieved by nding the average magnitude FEMrad over all directions and multiplying by 4r2 . The average over directions is ( |x x a|2 = 2 2 |a| , 3 (3.21)

hence the total power radiated in vacuo by a non-relativistic particle is Prad = q2 |a|2 . 60 c3 (3.22)

This is the Larmor formula. It describes the power radiated in electric dipole radiation by a non-relativistic particle. This is made explicit by rewriting the acceleration in terms of the dipole moment d(t) = q X(t).

3.2

Coherent radiation from a collection of partices

We can calculate the radiation from a collection of charges in a system of length scale L by calculating the sum of the electric and magnetic elds from each particle, and then calculating the resulting Poynting vector. When the system radiates coherently we show that a system comprised of N particles of charge q radiates like a macroparticle of charge N q . Because the power radiated is proportional to the charge, the power radiated in coherent emission scales as N 2. Suppose the system changes on a typical timescale . Then if the light travel time across the system is smaller than this timescale (L c) then there will be no appreciable changes from one side of the system to the other and dierences in retarded times across either end of the system can be neglected. However, if the system does change on a timescale the characteristic frequency of the radiation is roughly the inverse of this time (recall the power spectra from pulses of duration t in the previous Chapter) 1 . (3.23)

43

Now, in order for the phase dierences from all particles in the system to be small, we require that the size of this system be less than a wavelength: L c c = . (3.24)

In the non-relativistic limit the electric eld from this collection of particles is then E=
i

Ei ,

(3.25)

where Ei is dened by (3.13), and a similar equation holds for the magnetic eld. For radiation received at a point far outside the system dierences in the retarded time from radiation across the system are unimportant and the all the i vectors (i.e. all the Ei s) point in the same direction. Thus the dipole moment x due to the collection of particles is dcoherent sum =
i

qi xi ,

(3.26)

and the total power radiated is P = |dsum (t)|2 N 2 q2 a = . 60 c3 60 c3 (3.27)

Thus we see that, under the assumption of coherent emission, the total power radiated scales as the square of the number of radiating charges. Coherent emission is not the normal sort of radiation received from most large astrophysical sources. The most notable exceptions include radiation from pulsars and solar radio bursts. In both cases plasma processes are responsible for maintaining the coherence of the radiating system. When the radiating particles are suciently sparsely packed that there are signicant dierences in the retarded times between the elds emitted from the various particles, the radiation is incoherent. In this case the electric eld is an incoherent some of the elds from each of the radiating particles which is written in the form
N

E=
j

Ej =
j

E0 eij .

(3.28)

The phases j dier signicantly between the various particles, so that the electric eld is an incoherent sum of the electric (and magnetic) elds from all the particles. Now recall that the magnitude of the Poynting vector is proportional to E 2 . When we calculate the emission from the group of particles, we need to calculate E 2 which, in this case is (c.f. random walks on a 2-d plane)
2 E2 incoherent sum = N E0 .

(3.29)

44

It follows that the power radiated by a system of incoherently radiating particles is P = N q2 a N q2 60 c3 (3.30)

Thus we see that the total power radiated by the system is much smaller, as it is only linearly proportional to the number of radiating particles.

3.3

Spectrum

We can calculate the energy radiated per unit frequency by Fourier transforming the electric eld. Hence one has ( ) = E = But one has (see Ex. 2) (t) = d

1 2 1 2

dt eit E(t) dt eit (t) d sin . 40 rc (3.31)

( ) d, 2 eit d

(3.32)

so the Fourier-transformed electric eld is ( ) = sin E 40 rc 1 2


ei(

)t

dt

( ) d . 2 d

(3.33)

The quantity in curly brackets equals ( ) so the integrals reduce to ( ) = E 1 ( ) sin 2d 40 rc (3.34)

Thus the energy radiated per unit frequency is Urad =


2 dW 4 ( )|2 = q |F( )| , = 2 3 |d d 6 0 c 6 2 0 m 2 c3

(3.35)

where F( ) is the temporal Fourier transform of the force exerted on the particle.

3.4

The general multipole expansion

We derived the Larmor formula for emission from a single moving charge, but we would ideally like to calculate the emission from a distribution of charge and current. When we consider this more general situation we see that the Larmor formula merely represents the rst term in an expansion for the total emission from the system. It turns out that the total emission from the system may be 45

separated into an electric dipole term (i.e. the Larmor term), a magnetic dipole term, an electric quadrupole term and so on. Recall that the vector potential A depends on the current density J(t, x). If we do not specify the exact form of the current density but carry this term in our calculations we nd that the power radiated depends on J(, k), the Fourier transform of the current density. The expansion in the power radiated by the current density is related to an expansion of J(, k): J (, k) = = dt eit dt eit d3 x eikx J(t, x) 1 d3 x 1 ik x (k x)2 + . . . J(t, x). (3.36) 2

After a considerable amount of algebra we can re-write the components of this expansion in terms of multipole moments (i.e. d the electric dipole, the magnetic dipole, qij the magnetic quadrupole and so on). Using tensor notation, the i th component of J(, k) is Ji (, k) = = i di (t) + i[k (t)]i qis (t) + t 2 t 1 idi ( ) + i[k ( )]i ks qis ( ) + . 2 dt eit

(3.37)

These terms are associated with electric dipole, magnetic dipole and electric dipole emission respectively. It is conventional to treat emission by dierent multipole moments of the same system as independent processes. This is not strictly correct because the total current density is a sum over the individual currents from all the multipole moments. Thus the cross terms between multipole moments can, in principle, contribute to the total emitted power. However, in practice we are usually only interested in the emission due to the lowest multipole moment that gives a nonzero contribution to the radiated power. Thus interference with the cross terms is only a small correction if the expansion converges rapidly. Indeed, a detailed calculation shows the total power radiated in transverse waves due to the cross terms averages to zero.

3.5

Radiation Reaction

Classical electromagnetic theory does not explicitly incorporate the conservation of energy and momentum. Thus the back reaction of a particle when it emits a photon of radiation is not automatically taken into account. A proper treatment of the radiation reaction involves a quantum treatment, but in lieu of this we chose merely to supplement classical theory by postulating a radiation reaction force Freact (t) such that minus the work done by this force is equal to the power radiated: P (t) = v(t) Freact (t). 46 (3.38)

We can partially integrate the Larmor formula dt P (t) q2 (t)|2 dt |v 60 c3 q2 (t) dt v (t) v = 60 c3 =

(3.39)

to identify the radiation reaction force: Freact (t) = q2 (t). v 60 c3 (3.40)

One must be careful in interpreting the radiation reaction force. If interpreted as a real force it leads to unphysical solutions and acausal eects. The equation of motion of a particle subject to some other external force, F0 , say, in conjunction with the radiation reaction force is given by the Abraham-Lorentz equation of motion: (t)] = F0 (t) (t) 0 v m [v 0 = q2 60 c3 (3.41)

In the absence of an external force this has the physically acceptable solution (t) = 0, but it also has the physically unacceptable solution v (t) et/0 . v (3.42)

One way around this is to replace the Abraham-Lorentz equation with an integro-dierential equation

(t) = mv
0

dx ex F0 (t + 0 x).

(3.43)

This equation does not give rise to a runaway solution, but it does imply acausal eects. The problem is that (3.43) implies that the acceleration acting on a particle at time t is due to the exact value of the force F0 at the later time t + 0 x. Thus the particle is required to respond to a force before it has acted. However, this preacceleration is of no practical signicance because it occurs on a time scale that is intrinsically unmeasurable. All these diculties associated with the radiation reaction force are due to the fact that it is not strictly correct to simply equate Freact with the term appearing under the integral in equation (3.39). We are eectively equating terms that appear in two dierent integrals. The theory only requires that P (t)dt equal the work done by the radiation reaction force (another integral).

3.6

Example

What is the power and polarization emitted by a particle of mass m and charge q moving at speed v in a circle of radius r? 47

To solve this problem, we will consider the motion of the particle on the x-y plane whose position at time t is r(t) = r(cos t, sin t, 0) and its velocity and accelaration are (t) = r r(t) = r ( sin t, cos t, 0) = v ( sin t, cos t, 0) v ( cos t, sin t, 0). (3.45) (3.46) (3.44)

Now suppose we are looking at the radiation from the particle a distance R away at position x = (sin sin , sin cos , cos ) Recall the electric eld radiated by the particle is Erad = After some algebra we nd Erad = q v cos2 cos t + sin2 sin sin[ t], cos2 sin t + sin2 cos sin[ t], 40 Rc2 1 (3.49) cos[ t] sin 2 . 2 q [ x ( x a)]. 40 Rc2 (3.48) (3.47)

The time-averaged power received from the particle radiation is P (t)


t

= =

q2 E2 = 2 0 c 16 2 0 0 R2 c5 2 2 2 q v (1 + cos2 ). 32 2 0 R2 c3

v2 2 3 + cos 2 2 cos(2 2t) sin2 4 (3.50)

We compute the polarization by resolving the electric eld into two com (i.e. we want to resolve the ponents mutually orthogonal to the direction x oscilallations in the electric eld into local x and y values). For the purposes of simplicity we take x = (0, sin , cos ). Then this forms a mutually orthogonal set with the vectors ax = (1, 0, 0) and ay = (0, cos , sin ). One can then show from the denitions of the Stokes parameters that I U V = = = A(1 + cos2 ) A(1 cos ) 0 2A cos
2

(3.51) (3.52) (3.53) (3.54)

Q =

with A = (qv/16 2 0 R2 c3 ). By computing the total power in the polarization vector, Q2 + U 2 + V 2 , it follows that the radiation due to cyclotron emission is fully polarized. 48

3.7

Exercises

( ) = 2 q X( ). ( ) = 2 X( ) and hence that d 1. (a) Show that X (b) Show that (3.22) and (3.35) satisfy the relation

Erad =
0

d Urad ( ) =

dt P (t).

(3.55)

(Hint: Urad is an even function of . Use the power theorem.) 2. Consider the electromagnetic eld at a point described by polar co-ordinates r, due to a charge q in constant rectilinear motion at velocity c along the polar axis. Assume that the particle is at the origin r = 0 at time t = 0. (a) Show that the retarded time is given by the smaller of the solutions of the quadratic equation (t tr )2 = r2 /c2 2tr (r/c) cos + 2 t2 r. (3.56)

(b) Show that the electric and magnetic elds of a charge q in constant rectilinear motion at velocity c are given by E B = = x r q 2 40 (r x )3 q0 c r x , 4 2 (r x )3 (3.57) (3.58)

with = (1 2 )1/2 . (c) In the non-relativistic limit (10.23) gives the Coloumb eld. Give a physical interpretation of the non-relativistic limit of (10.24) in terms of the Biot-Savart law. 3. To lowest order, emission of gravitational radiation may be treated as quadrupole radiation. Estimate the energy given o in gravitational waves by a cyclist shaking his st at a passing motorist, given that the power emitted in gravitational waves is G d3 Qm P = 5 5c dt3
2

(3.59)

where Qm is the mass quadrupole moment. (Write the mass quadrupole moment in terms of the mass of the cyclists st and forearm, the length of the forearm and average period of each shake.) Use the fact that the muscles are almost totally nonconservative to estimate the fraction of energy emitted in gravitational radiation relative to the total energy expended.

49

Chapter 4

Radiation from Relativistic Systems


Reference: Rybicki & Lightman, Chapter 4
Here we review the basics of special relativity and consider aspects that are relevant to emission from relativistic systems.

4.1

Four-vector Notation

The basic four vector describes an event that occurs at time t and point x. The 4-vector for an event is written x = [ct, x] where the entries are the time and space components of the 4-vector respectively. 4-tensor indices are usually greek symbols which run over 0, 1, 2 and 3, with the time-like component being the zeroth component. Examples of 4-vectors are an event x = [ct, x] 4-velocity u = [c, v] 4-momentum p = [/c, p] wave 4-vector k = [/c, k] 4-current density J = [c, J] 4-potential A = [/c, A] 4-force F = [ v F/c, F], where = 1/ 1 v 2 /c2 is the Lorentz factor and = mc2 is the energy. Lorentz transformations convert 4-vectors between inertial frames. A Lorentz transformation of a 4-vector b from one inertial frame K to another inertial frame K is written b = L b .

(4.1)

The particle is stationary in the rest frame K , while it is observed to move toward increasing z from the point of view of an observer in the laboratory frame K . We can write this velocity of K relative to K as v = c with 50

The reverse Lorentz transformation is found by making the replacement . The matrix above describes what is referred to as a boost along the z axis. One can determine the boost along any arbitrary axis by rotating the three spatial co-ordinates into a frame in which the boost occurs along the z -axis. A quantity that is unchanged under a Lorentz transformation is called an invariant. In 4-tensor notation the invariant formed from any two 4-vectors plays a role analogous to the scalar product of two 3-vectors in 3-tensor notation. The invariant formed from the 4-vectors a = [a0 , a] and b = [b0 , b] is a0 b0 a b. To take account of the minus sign one denes the contravariant components a with raised index and the covariant components a = [a0 , a] with lowered index for each 4-vector. Covariant components transform like the 4-gradient: = /x = [/ct, / x]. The invariant formed from a and b is a b = a b = a0 b0 a b. (4.4) (4.3)

= (1 2 )1/2 . The transformation matrix for the Lorentz transformation from the frame K into K where v is along z -axis is 0 0 0 1 0 0 . (4.2) L = 0 0 1 0 0 0

An invariant of particular importance is the proper time , which is interpreted as the actual time according to an observer moving in the rest frame. Consider the trajectory of a particle written in covariant form as a function of the proper time x = X ( ). (4.5)

The time component X 0 ( ) is such that the invariant dx dx = c2 d 2 formed from the dierential of (4.5) denes an element d of proper time. This gives d = dt , (4.6)

where is regarded as a function of t in a specic inertial frame, and integrating (4.6) along the particle trajectory gives the proper time. The 4-velocity and 4acceleration are given by dierentiating with respect to proper time: u = dX ( )/d, a ( ) = d2 X ( )/d 2 . (4.7)

The covariant form of Newtons equation of motion is ma ( ) = F . (4.8)

51

4.2

Specic Examples of Lorentz Transformations

Let us use the formalism introduced in the previous section to discuss the special relativistic eects of length contraction and time dilation. Length contraction Consider a rod at rest in frame K that moves past an observer at velocity v = (0, 0, vx ). Consider two points in the rods frame r 1 and r2 , located along a rod. The distance between these two points depends on the inertial frame in which it is measured. We employ the Lorentz matrix introduced in the previous section to derive the positions of the two points along the rod in the observers frame, K :
2 t 1 = (t vz rz 1 /c ) r1 x = r1 x r1 y = r1 y r1 z = (r1 z vz t) 2 t 2 = (t vz rz 2 /c ) r2 x = r2 x r2 y = r2 y r2 z = (r2 z vz t).

(4.9)

Thus if the rod at rest has a length L = r2 z r1 z in its rest frame then it is measured to have a length r2 z r1 z = (r2 z r1 z )/ = L/ in the laboratory frame K . Hence we conclude that moving objects are shortened along their direction of motion; this is called Lorentz contraction.

Time Dilation A related eect is time dilation, whereby moving clocks are observed to run slow. Consider a clock at rest in frame K which is measured at times t 1 = 0 , 0] in K ). An observer in the , 0] and x = [ ct (i.e. 4-vectors x = [ ct and t 2 1 2 2 1 clock rest frame measures an interval t = t t between the two times. We 2 1 employ the Lorentz transformation described by (4.2) to relate the 4-vectors in K to 4-vectors in the laboratory frame K . One has
x 2

x 1

= [c (t2 vz z2 /c ), x, y, (z2 vz t2 )]

= [c (t1 vz z1 /c2 ), x, y, (z1 vz t1 )]


2

(4.10) (4.11) (4.12)

Hence the interval t is


2 2 t = t 2 t1 = [(t2 vz z1 /c ) (t1 vz z2 /c )]

(4.13)

Now if the clock is stationary in K its z position remains constant so we have z1 vz t1 = z2 vz t2 . This implies z2 z1 = vz (t2 t1 ) so (4.13) simplies to t = =
2 2 (t2 t1 ) + vz /c (t2 t1 ) 1 2 (t2 t1 ) + 2 2

52

t2 t1 ,

(4.14)

where we have used the relation 2 + 2 2 = 1. This result leads us to the conclusion that t = t hence that moving clocks run slow by a factor . Velocity transformation There are two ways to derive the velocity transformation between intertial frames K and K . One can take the derivative of x, y and z with respect to t, one can also use the transform the velocity 4-vector u = [c, v] using (4.2). Either way, we nd ux = dx u x = vu dt 1 + c2z u dy y = vu dt 1 + c2z dz u + v = z vu , dt 1 + c2z (4.16) (4.15)

uy uz

where v is the speed of the frame K relative to K . We can simplify these expressions by decomposing the velocities into components parallel and perpendicular to the frame velocity v: u u = = u + v 1+
vu c2

u 1+
vu c2

(4.17)

So if the direction of an objects motion in frame K is tan = then its direction in frame K is tan = u = u 1+ u 1+
vu c2 vu c2

u u

(4.18)

u + v

u sin . (u cos + v )

(4.19)

53

Beaming and the special relativistic Doppler eect Dopeler boosting the tendency of stupid ideas to appear brighter when they come at you rapidly. The transformation of velocities has important implications when considering radiation from particles travelling at relativistic velocities. Consider radiation emitted at an angle from a particle travelling at v/c 1. Then the radiation will appear to be emitted at a dierent angle in the frame K : tan = sin . (cos + v/c) (4.20)

Inspection of (4.20) shows that the emission from all angles is swept into a forward cone of opening angle 1/ , except from a small range of angles 1/ behind the particles direction of motion. Relativistic motion also aects the apparent frequency of the radiation. We can calculate this eect by transforming the wave 4-vector k = [/c, k] using (4.2). We nd kx
ky kz

= kz vz = kx = ky = (kz vz /c). (4.21)

Now, in a vacuum one has k = /c, so the rst relation can be written as the relativistic Doppler relation = (1
v c

cos )

(4.22)

where is the usual direction of emission relative to the velocity of frame K . Note that there are two separate physical eects embodied in this equation. The factor of on the denominator is an inherently relativistic phenonemon due to the time dilation of events occuring in frame K and observed in the laboratory frame K . The second term in the demoninator represents the classical Doppler eect, that the frequency of waves (e.g. even sound waves) are altered as the motion of the source approaches the wave velocity.

4.3

Electromagnetism

Maxwells equations are Lorentz covariant, which is to say that they hold in any inertial frame. In other words, electromagnetism is already consistent with special relativity. However, the electric and magnetic elds are not Lorentz covariant. To see this, imagine an observer who measures the elds due to a stationary charge: this particle has an electric eld but no magnetic eld, since 54

the charge is not moving (i.e. = 0, J = 0). However, an observer moving relative to this charge discerns this as a moving charge and thus as a current (i.e. = 0, J = 0) and measures both an electric eld and a magnetic eld. We can determine how electric and magnetic elds transform between inertial frames from the Lorentz transformation of the 4-potential A = [/c, A]. If we do this, we see that it proves useful to dene the antisymmetric tensor 0 Ex Ey Ez Ex 0 Bz By . (4.23) F = Ey Bz 0 Bx Ez By Bx 0 This tensor transforms in the usual way
F , L F =L

(4.24)

with = L , L where is the Minkowski metric = (4.25)

After some algebra (exercise!) we nd that the electric and magnetic elds transform as E = E B = B E = (E + B) B = (B E). (4.27)

1 0 = 0 0

0 1 0 0

0 0 1 0

0 0 . 0 1

(4.26)

Thus we see that the components of the electric and magnetic elds parallel to the direction of motion remain the same, but their components perpendicular to the direction of motion do change. Because F is Lorentz invariant, various quantities derived from it are also Lorentz invariant. In particular the quantities F F detF = 2(B2 E2 ), = (E B)2 . and (4.28) (4.29)

are Lorentz invariant, which is to say that B 2 E 2 = B 2 E 2 and EB = E B . The eld of a uniformly moving line of charge Here we attempt to gain an intuitive appreciation of the fact that Maxwells equations are Lorentz covariant but that the electric and magnetic elds themselves are not. To this end, we consider the elds and forces due to a line of moving charge. 55

Let us take a line of positive charges moving along a line with speed v to the right. We assume that there are suciently many charges that we can consider them as a continuous line charge (coloumbs per unit metre). Superimposed on this string of positive charges is a negative one with charge density proceeding along the line with speed v . Thus, we have a net current to the right of magnitude I = 2v. (4.30)

This wire is at rest in the laboratory frame K and in this frame the wires are electrically neutral since the charge densities are equal and opposite + () = 0. Now consider a test charge q a distance r away from the wire that is itself moving at a speed u to the right. Consider this system in the rest frame of the test charge, K . By the Einstein velocity addition rule, the velocities of the positive and negative lines of charge are now v = vu . 1 vu/c2 (4.31)

Because v is greater than v+ , the Lorentz contraction of the spacing between negative charges is more severe than that between the positive charges: the magnitude of the current density in the negatively charged wire is higher than that in the positively charged wire. Thus, in the rest frame of the test charge, the wire carries a net negative charge! The new charge densities are = , where = (4.32) 1
2 /c2 1 v

(4.33)

You can derive this result by noting the form of the 4-vector for the current density and applying a Lorentz transformation to this. Inserting the expressions for v , we nd = and the net charge in the wires is tot = + = (+ ) = c2 2uv . 1 u2 /c2 (4.35) 1 uv/c2 , 1 u2 /c2 (4.34)

Thus, the unequal Lorentz contraction of the positive and negative current carrying lines implies that the wires which are electrically neutral in one frame will be charged in another. We can calculate the electrical force on the test charge in the frame K . Recall that a line charge tot sets up an electric eld E= tot , 20 r 56 (4.36)

so our test charge experiences a force in the frame K of F = qE = v 0 c2 r qu 1 u2 /c2 . (4.37)

But if there is a force in K there must also be a force in the rest frame of the wire, K . We can determine this force by Lorentz transforming K , for which we nd F = 1 u2 /c2 F = v qu . 0 c2 r (4.38)

So we conclude that the charge is attracted to the wire by a force which is electrical in K but distinctly non-electrical in the frame K , for which the net charge density is zero. Thus we are led to conclude that, taken together, electrostatics and relativity imply the existence of another force, which is, of course, magnetic. Writing the current as 2v as the current I this force is F = 0 I 2r qu (4.39)

This is just the magnetic force due to a long straight wire of current, and the term in brackets is the magnitude of the magnetic eld!

4.4

Emission from Relativistic Particles The Relativistic Larmor formula

Let us return to the Larmor formula for the electric dipole radiation emitted by a moving charge P = q 2 |a(t)|2 . 60 c3 (4.40)

which imply = c In the rest frame we have v = 0 with = 1 and a = v 2 2 u = [c, 0], a = [0, c] and a a = c . Hence the Larmor formula is written in the form P = q 2 a a , 60 c3 (4.41)

in the rest frame. The left hand side is the ratio of the time-like components of two 4-vectors (the energy radiated and the time) and hence is an invariant. The right hand side is already in an invariant form. Hence the special theory of relativity implies that (4.41) holds in all inertial frames. We can write the relativistic Larmor formula explicitly using the relations a a a = = , ( ) + (1 2 ) ], c 4 [ | |2 ]. c2 6 [ 57
2

(4.42) (4.43)

Thus the generalization of the Larmor formula is written in full as | |2 q2 P = . 60 c (1 2 )3


2

(4.44)

One can resolve the acceleration into components parellel and perpendicular to the direction of instantaneous motion in the rest frame of the particle: a = 2 3 a , a = a , where is the Lorentz factor of the particle in the laboratory frame. The power is then P = q2 q2 2 2 a a = 4 (a2 + a ). 60 c 60 c (4.45)

In terms of the force acting on the particle this is P = q 2 2 (|F|2 | F|2 ) . 60 c m 2 c2 (4.46)

4.5

Further Invariant quantities

Consider a group of particles which occupy a region of space [x , x + d3 x ] and possess momenta [p , p + d3 p ] in a frame, K comoving with the particles (i.e. their rest frame). To rst order is no spread in energy (d = 0) since the energy is quadratic in the momentum. We can say that the particles occupy an element of postion-momentum phase space dV = d3 p d3 x . (4.47)

Now recall that for motion along the x-axis the spatial element d3 x transforms into the laboratory frame K like d3 x = 1 d3 x , (4.48)

since dx = 1 dx, dy = dy and dz = dz . The element of momentum transforms like d3 p = d3 p , (4.49)

since dp y dpz = dpy dpz and dpx = (dpx + d ) = dpx . Thus we see that the quantity dV is a Lorentz invariant:

dV = dV .

(4.50)

It also follows that the number density f = dN/dV of particles within the phase space element dV is also Lorentz invariant because the number of particles within dV is a countable quantity and thus Lorentz invariant. We can relate the number density of photons, f , within a phase volume to the specic intensity of the radiation, I . Now the energy density of the 58

radiation per unit solid angle per unit frequency u () (see Chapter 1) can be written u () d d = h f p2 dp d. (4.51)

We can write this in a more useful form using the fact u () = I /c and p = h/c to nd I = Lorentz invariant. 3 (4.52)

This relation proves extremely useful in many instances in relativistic astrophysics. From this relation, we can also argue that the follow quantities are also Lorentz invariant: S / 3 , and j / 2 .

4.6

Exercises

1. What is the appropriate form of the acceleration 4-vector a so that the Lorentz transformation formula (4.1) with the transformation matrix (4.2) correctly transforms this 4-vector between inertial frames. [Hint: rst work out how accelerations transform between inertial frames (see Rybicki & Lightman ex. 4.3).]

59

Chapter 5

Bremsstrahlung
Reference: Rybicki & Lightman, Chapter 5 Melrose & McPhedran, Chapter 21
Bremsstrahlung is variously used to refer to (i) emission due to an accelerated charge and (ii) such emission when the acceleration is due to the Coulomb eld of another particle. Here we consider bremsstrahlung in the context of the second, more restrictive sense. This is often referred to as free-free emission in the astrophysical literature. We also consider the corresponding absorption process, called collisional damping, or free-free absorption.

5.1

Qualitative Discussion of Bremsstrahlung

Bremsstrahlung arises due to Coulomb interactions in a plasma. Only encounters between dierent species specically those between electrons and ions need be considered. Encounters between like species yield no dipole emission, as the second time derivative of their dipole moments is easily shown to be zero. In our treatment of bremsstrahlung we treat Coulomb interactions as encounters between pairs of particles. We regard the encounters as perturbations on the orbit of a test charge, with the unperturbed orbit taken to be constant rectilinear motion. The eect of each interaction is treated as a separate encounter. It is not obvious that we are justied in treating collisions as separate binary encounters. The Coulomb eld of each particle falls o as 1/r2 but the number of particles in a shell between r and r + dr increases as r2 dr, so that interactions at all distances appear to be equally important. In practice, the eld suciently far from the test charge contains roughly equal contributions from both negative and positive species, so that the mean eld due to particles at large distances does not inuence the test charge. In plasma physics this concept is known as

60

Debye shielding. The eld is modied from (x) = q/40 r to (x) = qer/D , 40 r (5.1)

where D is the Debye radius. The exponential cuto in the eld implies that only particles within the Debye sphere inuence the test particle. Nevertheless, this number is considerable, and in laboratory and astrophysical plasmas the Coulomb logarithm the natural logarithm of the number of particles in the Debye sphere is between 10 and 20. Our treatment of bremsstrahlung proceeds by rst considering the electric dipole emission associated with a test charge. We then consider the orbit of an electron due to interactions with a single ion in the plasma; here we make the straight line approximation in which we assume that the unperturbed orbit of the electron is a straight line. This is the basis for treating multiple encounters in the plasma. We consider the specic case when the velocity distribution is thermal. However it is not obvious from our treatment how to set the upper and lower bounds of the integration over all encounters. These bounds may be extracted from a quantum mechanical considerations which we not discuss here; instead we introduce correction factors, known as Gaunt factors, to the expressions dervied from our classical treatment. We then consider the corresponding absorption process of free-free emission, called collision damping, which modies the observed spectrum of a free-free emitting source at the low frequency end. We conclude with a discussion of the free-free emission due to a relativistic distribution of particles.

Figure 5.1: The Coulomb interaction between an electron with an ion generates bremsstrahlung emission.

5.2

The Straight Line Approximation

We are now in a position to consider the radiation from an electron in a plasma moving in the Coulomb eld of a nearby ion, as depicted in Figure 1.1. The 61

properties of the radiation are derived by determining the orbit of the electron. The straight line approximation is a perturbation treatement in which the zeroth order orbit of an electron moving past an ion, which may be considered immobile, is taken to be rectilinear motion: X(0) (t) = b + vt, (5.2)

where b is the point of closest approach to the ion, and the velocity is assumed to be constant to zeroth order. For an ion of charge Ze the rst order acceleration is then
2 (0) (t) = Ze X (t) . F(1) (t) = me X 40 |X(0) (t)|3

(5.3)

This acceleration may then be used to derive higher-order corrections to the orbit, however, this is not necessary for our purposes. The power radiated is given by substituting into the Larmor formula P (t) = e2 60 c3 Ze2 40 me |X(0) (t)|2
2

(5.4)

and the total energy radiated per encounter is found by integrating the power over all time:

Erad =

P (t) dt

= =

2 Z 2 e6 3 3(40 )3 m2 ec

(b2

dt + v 2 t2 )2 (5.5)

Z 2 e6 . 3 3 3(40 )3 m2 e c vb

The spectrum of the radiated energy is derived by substitution into (3.35): Urad ( ) = e2 F(1) 3 6 2 0 m2 ec
2

(5.6)

An explicit form of the spectrum requires that we nd the temporal Fourier transform of the force experienced by the electron. The integral is evaluated in terms of Bessel functions of the second kind, Kn (x) also known as Macdonald functions and yields F(1) ( ) = = Ze2 40

dt eit

b + vt (b2 + v 2 t2 )3/2 v i K0 v b v .

(5.7) (5.8)

2Ze2 b K1 40 v 2 b

b v

The energy radiated per unit frequency then reduces to Urad ( ) = 8 Z 2 e6 2 2 K1 3 4 3 (40 )3 m2 ec v 62 b v
2 + K0

b v

(5.9)

The limiting forms of the Macdonald functions for small and large argument yield the following approximate spectrum for the bremsstrahlung due to a single electron-ion encounter with impact parameter b: 8 Z 2 e6 2 Urad ( ) 3 4 3 (40 )3 m2 ec v
v2 b2 2 v 2b

exp

2b v

b v b v

1 . 1

(5.10)

Thus the energy per unit unit frequency is at at frequencies lower than v/b and falls o rapidly at frequencies larger than v/b. This type of spectrum is characteristic of emission due to an impulsive event of duration t = b/v : the frequency spectrum is white noise for (t)1 and falls of rapidly for (t)1 .

5.3

Multiple encounters

In practice an electron interacts with a large number of ions simultaneously. The electron emits continuously due to all the interactions with ions within a Debye radius. However, the resulting radiation from each single encounter is independent of the radiation from each other encounter; there is no interference due to the random positions and motions of the ions. Thus the total emission due to a single electron is determined by summing over all its interactions with ions in the plasma. Suppose that the number density of ions with charge Zi e is ni . Now the number of ions with speed v transversing the cross sectional area between impact parameters of radius b and b + db per unit time is 2b db ni v . The factor 2b db is the cross-sectional area of impact parameter space in a cylindrical geometry. The power radiated by a single electron is found by summing over ion species and integrating over impact parameters:
b2

P ( ) =
i

2ni v
b1

bUrad ( ).

(5.11)

We have introduced a lower limit to the range of impact parameters because the integral diverges as b1 0. The power also diverges as 0 if we take b2 as the upper cuto. This diculty is resolved by taking Debye shielding into account, which tells us that the Coulomb eld used to derive Urad in the previous section is only valid for distances less than the Debye radius, D . We return to the discussion of appropriate expressions for the lower cuto, b1 , in the next section. The integral in (5.11) is evaluated using the result
z1 2 2 z K0 (z ) + K1 (z ) = z1 K0 (z1 )K1 (z1 )

(5.12)

to obtain P ( ) =
i 2 Zi ni e 6 b1 16 K0 3v 3 (40 )3 m2 c v e

b1 v 63

K1

b1 v

b2 K0 v

b2 v

K1

b2 v

(5.13) This has low and high frequency limits P ( ) =


i 2 16 Zi ni e 6 3 3 (40 )3 m2 ec v

ln
2

2 b2 b1

exp

1 2bv

, v/b1

v/b1

(5.14)

The spectrum is essentially at at low frequencies, with the only dependence on being due to the slowly-varying logarithmic term. For v/b1 the amplitude of the spectrum is insensitive to the actual value of the minimum impact parameter, since it also only exhibits a logarithmic dependence on b1 . The spectrum cuts-o at a high frequency dictated by the minimum impact parameter, at v/b1 . The amplitude of the spectrum in this regime is independent of the maximum impact parameter, b2 .

5.4

Gaunt factors

We have used several approximations in deriving (5.13). There are diculties of nding appropriate upper and lower bounds to the integral over impact parameters; we have not even specied an appropriate lower bound yet. Our treatment of electron trajectories in the plasma in 5.2 was also approximate. Furthermore, the classical treatment does not take into account quantum mechanical eects when the energy of emission h is comparable to the electron kinetic energy me v 2 /2. It is convenient to combine all these corrections into the a single number called the Gaunt factor. One makes the correction by replacing the term ln(2/ b2 /b1 ) in (5.13) with the Gaunt factor as follows: 2 b2 3 . (5.15) ln G(v, ) = b1 The factor of 3/ arises as an historical convention.

5.4.1

Classical corrections

An obvious classical correction to (5.13) involves an appropriate choice of inner cuto, b1 . Since the dependence on b1 is logarithmic, it suces to obtain only an approximate value for thus cut-o. An obvious choice is the minimum impact parameter for which a deection through an angle of /2 occurs as an electron encounters an ion. This occurs for b1 = e2 Zi /(40 me v 2 ). An appropriate upper limit corresponds to the value at which the bremsstrahlung spectrum turns over: b2 = v/ . This yields a Gaunt factor 2 40 me v 3 3 G(v, ) = . (5.16) ln Zi e2 This factor is derived from purely classical considerations and does not taken into account quantum mechanical eects. 64

5.4.2

Quantum mechanical corrections

Quantum mechanical eects become important when the de Broglie wavelength of the electron, h /p, becomes comparable to the impact parameter. The appropriate cut-o b1 is the greater of either the impact parameter for which deection through /2 occurs or the de Broglie wavelength. In this latter case, the appropriate Gaunt factor is 3 2 me v 2 G(v, ) = . (5.17) ln h In this case the quantum mechanical expression (5.17) is used instead of the classical one in (5.16). Quantum mechanical eects are also important at high frequencies when the energy of the emitted photon is comparable to the energy of the electron. Indeed, the requirement of energy conservation is not built in to the classical treatment. This limitation is overcome by introducing the Gaunt factor v + v 3 , (5.18) ln G(v, ) = v v where the nal kinetic energy of the electron is 1 1 me v 2 = me v 2 h. 2 2 (5.19)

The Gaunt factor given by (5.18) reduces to (5.17) in the limit h me v 2 /2. It is possible to perform a full quantum mechanical derivation of bremsstrahlung involving interactions between the waveeld of the electron in the Coulomb eld of the ions. We do not state the results here but merely mention that emission at frequencies such that h > me v 2 /2 is still possible via free-bound transitions. The large density of bound states just separating bound and free transitions 1 makes bremsstrahlung ecient at energies near h 2 me v 2 .

5.4.3

Bremsstrahlung in a Thermal Plasma

Our expressioni for the emission due to bremsstrahlung in equation (5.13)does not include an average over the velocity of the particles. This average is performed once the distribution of electron speeds is specied. An obvious choice is when the electrons speeds obey a thermal distribution: fe (v) = 1 (2 )3 Ve6 exp |v|2 2Ve2 (5.20)

where we identify Ve = kT /me as the thermal speed of the electrons. The average over velocity in (5.13) requires that we average the quantity 1 b1 K0 v v b1 v K1 b1 v b2 K0 v b2 v K1 b2 v (5.21)

65

over this distribution. This average is complicated, but we can write the power radiated per thermal electron per unit frequency as P ( ) =
i 2 16 Zi ni e 6 3 3 (40 )3 m2 ec

2 1 Gth (Ve , ), Ve 3

(5.22)

where we have performed the important average over 1/v , 1 v = dv 1 fe (v) = v 2 1 , Ve (5.23)

and bundled the rest of the average over velocity as a correction into the Gaunt factor. This Gaunt factor is Gth (Ve , ) = K0 ( (5.24) h/me Ve2 ) exp h/me Ve2 3 ln(2me Ve2 / h ), h me Ve2 (5.25) . 2 2 h me Ve2 me Ve /2 h exp h/me Ve , The bremsstrahlung spectrum for a thermal distribution is therefore at for energies h me Ve2 and falls o expontentially at higher frequencies, as photon energy becomes comparable to the thermal energy of the electrons.

5.5

Collisional Damping

The absorption process corresponding to free-free emission is collisional damping, or free-free absorption. Collisional damping is important in certain astrophysical contexts, such as the radio emission from HII regions and the X-ray emission from galaxy clusters. In the special case of a thermal electron distribution it is possible to use Kirchhos law to treat this process. Recall that for a thermal distribution Kirchhos law relates the absorption coecient to the emission coecient and the specic intensity as follows: j = 4 B (T ). The specic intensity for thermal radiation is B (T ) = 2h 3 c2 eh/kT 1 = (2 )2 c2 2 h 3 . 2 h e /me Ve 1 (5.27) (5.26)

Since the radiation due to the many electron-ion encounters is isotropic, so is the emission, so the emissivity per unit volume is j = ne P ( )/4 with P ( ) given by (5.22) with (5.24). Inserting this into (5.26) with (5.27) yields =
i

ne

2 Zi ni e 6 2 3 (40 )3 m2 h 3 c e

2 1 K0 Ve 66

h me Ve2

/me Ve . (5.28) 1 eh

For energies h me Ve3 the argument of the exponential term is small and the term in brackets is approximately h /me V22 , so the absorption scales as 2 . 2 In the opposite limit, h me Ve , the exponential term is negligible and the absorption scales as 3 .

5.6

Other Considerations

An important point related to bremsstrahlung radiation from a formal point of view is that the emission necessarily occurs in a plasma, whereas our simple treatment has assumed the emission to occur in vacuo. We have thus disregarded the fact that the refractive index of the medium deviates from unity. In practice this is not a serious limitation. Consideration of this eect shows that one should write Pcorrect ( ) = n( )P ( ) (5.29)

a factor of such as (5.13). The refractive index of the plasma considered here is, to rst order, n( ) =
2 / 2 , 1 p

(5.30)

2 where p = e2 ne /0 me is the square of the plasma frequency. This correction is usually negligible for most plasmas. Most observations occur at angular frequencies well above p , where the refractive index is well-approximated as unity. Unlike the emissivity, the absorption coecent is independent of the refractive index. As a nal remark we note that the absorption coecient for bremsstrahlung cannot be negative, so that maser action due to bremsstrahlung is not possible.

5.7

Questions

1. Show that there is no electric dipole emission associated with bremsstrahlung due to electron-electron encounters. 2. Complete the derivation of (5.9) from (5.6) using the following relations for Macdonald functions: Kn u(xz ) = and
K 1 (z ) + K +1 (z ) = 2K (z ), 2 K 1 (z ) K +1 (z ) = Knu (z ). z ( + 1 2 )(2z ) 1 2x ( 2 )

dt

eixt , (t2 + z 2 ) +1/2

(5.31)

(5.32) (5.33)

67

3. Bremsstrahlung due to electron-electron collisions has no electric dipole component. The quadrupolar component of the emission is considered in the following exercise. The power radiated in a single encounter between two electrons is P =
2 v 2 + 11v e6 . 2 5 480 3 3 r4 0m c

(5.34)

a) Assume that the speed at innity is v0 and that the impact parameter is b, and argue that the dependence of the total speed v and the azimuthal component v of the velocity on radial distance r are determined by
2 v 2 = v0

e2 , 20 mr

v =

bv0 . r

(5.35)

b) Substitute (5.35) into (5.34) and derive an integral expression for the energy radiated in a single encounter as an integral over r with, r = (v 2 v )1/2 = 1 2
2 v0

e2 b2 v 2 20 20 mr r

(5.36)

c) Proceeding by analogy with (5.11), with v replaced by v0 , write down a formula for the power radiated by a single electron due to continuous encounters with surrounding electrons as an integral over impact parameter. d) Carry out the integrals over b and r, noting that the limits on the range are determined by r = 0, v . Hence show that the mean power radiated per electron in electron-electron bremsstrahlung is given by P =
4 ne e4 v0 . 2 360 mc5

(5.37)

[Hints: Reduce the double integral to the form 0 dr (r L)3/2 /r7/2 , with 2 L = e2 /20 mv0 . Change variables by writing r = L(1 + tan2 ).] The emission due to electron-electron encounters is weaker than that due to electron-ion encounters by a numerical factor times v 2 /c2 , where v is the speed of the emitting electron. It follows that electron-electron bremsstrahlung is unimportant in non-relativistic plasmas but becomes signicant at suciently high temperatures. 4. Show that the Gaunt factor given by (5.18) reduces to (5.17) in the limit h me v 2 /2. 5. For a radioastronomical source of bremsstrahlung that is optically thin and is spatially resolved, the emission is characterised by the temperature and quantity called the emission measure. a) Show that on integrating the power radiated per unit volume per unit 68

frequency and per unit solid angle ne P ( )/4 , with P ( ) given by (5.22) through a source of thickness L along the line of sight, the specic intensity of bremsstrahlung implies a brightness temperature
3 2 3 TB c c 32 2 r0 G(Ve , ), = EM Te 3 2 2 Ve 3

(5.38)

with r0 = e2 /40 mc2 the classical electron radius and where the emission measure for bremsstrahlung is dened by
L

EM =
0

dz
i

2 Zi ni ne .

(5.39)

b) An HII region is a sphere of fully ionised plasma around a bright star. A model for an HII region implies EM = 104 pc cm6 and Te = 104 K. Estimate the expected radio brightness temperature TB .

69

Chapter 6

Synchrotron Emission Theory


Reference: Melrose & McPhedran, Chapter 24 Rybicki & Lightman, Chapter 6 We also acknowledge D.B. Melrose, whose lecture notes this chapter and the next draw from substantially.
Synchrotron emission refers to radiation by ultra-relativistic particles as they orbit magnetic eld lines. It is the dominant emission mechanism in radio sources, and it is important in the consideration of emission at other wavelengths. This is the case for active galactic nuclei, in which synchrotron emission is also important at optical and even X-ray wavelengths via, for instance, induced Compton scattering. We rst discuss the radiation pattern due to an ultra-relativistic particle without regard to any specic emission process. We derive formulae for the emission due to the synchrotron radiation from a single electron, before examining the astrophysically important case in which the emission is due to a power-law distribution of electron energies. We then discuss the corresponding absorption process, known as synchrotron self-absorption. The chapter concludes with a brief look at the cyclotron emission, which is gyromagnetic emission from non-relativistic particles.

6.1

Emission by Relativistic Particles

Many features of the emission of highly relativistic particles can be derived from special relativity without reference to the specic emission mechanism involved. Suppose we have a highly relativistic particle whose instantaneous velocity is v along the z axis, as observed in the laboratory frame, K . We now dene another the frame K0 in which the particle is momentarily at rest. The frame K0 , as 70

viewed from K , is therefore moving along the z axis at velocity v, as illustrated in Figure 6.1, and the frame K , as viewed from K0 , is moving along the z0 axis at velocity v. In the frame K0 the particle is momentarily at rest. The radiation pattern in the rest frame K0 is dipolar with the the dipole axis along the direction of the instantaneous acceleration in K0 . The actual shape of the radiation pattern is unimportant for present purposes; it merely suces that the pattern is not highly anisotropic.
x0 K0 - particle rest frame x K - laboratory frame

P -v z0

Figure 6.1: The laboratory frame K moves at velocity v relative to the rest frame K0 . A particle at rest in K0 has velocity v in K . Consider a plane wave in the rest frame, K0 , with wavenumber k0 and angular frequency 0 propagating at an angle 0 to the z0 axis. Applying a Lorentz transformation, the wave parameters , k , in K are related to those in K0 by = (0 + k0 v cos 0 ), 0 = ( kv cos ), k sin = k0 sin 0 , k0 sin 0 = k sin , k cos = (k0 cos 0 + 0 v/c2 ), k0 cos 0 = (k cos v/c2 ),

(6.1)

with = (1 2 )1/2 , = v/c. Recall that the quantity |k|2 c2 2 = 0 is a Lorentz invariant, hence one has k = /c, k0 = 0 /c, and (6.1) implies = (1 + cos 0 ), 0 cos = cos 0 + , 1 + cos 0 sin = sin 0 . (6.2) (1 + cos 0 )

For a highly relativistic particle one assumes 1 and expands in powers of 1 : 1 1/2 2 , (6.3)

so that diers from unity by a term of order 2 . We now argue that most directions of emission in the rest frame transform into forward directions as observed in the frame K . The third equation of (6.2) implies that sin is 71

always of order 1/ provided that 1 + cos 0 and sin 0 are of order unity. The proviso is satised except for a small range of angles 0 of order 1 . This leads to the following property of the relativistic emission: All angles of emission in K0 , transform into a forward cone with halfangle of order 1 about the direction v in K , with the exception of a range of order 1 about the direction v. Thus all the emission is swept into a narrow forward cone except for a small amount from the side corresponding to the reverse direction of propagation . A similar argument can be made for the energy of the emission. Inspection of (6.2) shows that the frequency in K is of order times the frequency 0 in K0 , except for the small range of 0 of order 1 around cos 0 = 1. It follows that nearly all the photons emitted in K are conned to the forward cone and have an energy of order times the energy of the corresponding photon in K0 . Thus: All but a fraction 1 of the power emitted in K is conned to a forward cone with half-angle of order 1 about the direction v. For gyromagnetic emission this implies that all but a fraction 1 of the power emitted by a particle with pitch angle is conned to angles satisfying = + O( 1 ). (6.4)

Thus we are led to the following picture for an ultra-relativistic particle orbiting a magnetic eld line. As a particle spirals about a magnetic eld line, it traces out a spiral with a pitch angle dened by the components of its velocity perpendicular and parallel to the eld line. The particle radiates as it is continuously accelerated by the magnetic eld. All but a fraction 1 of the power is emitted on the surface of a cone with half-angle = and thickness of order 1 (see Figure 6.2). It is useful to picture the radiation pattern in K0 as forward-sweeping lighthouse with a narrow beam that continuously traces out a cone dened by the particles pitch angle. If the particle pitch angle is small most of the emission is beamed nearly along the direction of the magnetic eld line. For pitch angles /2 the emission is beamed perpendicular to the eld lines.

6.2

Semiquantitative Treatment

The properties of synchrotron emission may be derived in a semiquantitative manner by combining some exact results for gyromagnetic emission and the foregoing arguments concerning emission by relativistic particles. One exact result from the classical electromagnetic theory is for the power radiated by an accelerated charge in vacuo. We can derive this from the Larmor formula. The position of a particle spiraling in a magnetic eld along the z -axis may be written X(t) = R0 (cos e t, sin e t, v t) 72 (6.5)

1/

Figure 6.2: The emission pattern of a relativistic particle with 1.

Figure 6.3: The emission pattern of a gyrating particle with 1. The beam sweeps around in a circle as the particle gyrates. The emission occurs roughly on the surface of a cone with opening angle = with R0 = v sin , e e = eB , me (6.6)

where m is the mass of the particle. The acceleration in the rest frame of the particle is a = e v sin . Transforming this into the frame moving in the direction of the particle at speed v gives (see Qu. 1) a = 2 v e = 2 v sin Substitution into the Larmor formula yields P = e2 2 2 2 sin2 . 60 c e (6.8) eB me (6.7)

The emission in gyromagnetic radiation actually occurs due to a resonant interaction, and so must occur at a sequence of harmonics, s. The emission 73

must satisfy the resonance condition, k v s = 0 , (6.9)

where k and v are parallel the components of k and v relative to the magnetic eld. From a classical viewpoint, (10.34) corresponds to the wave frequency being an integral multiple of the gyrofrequency in the inertial frame in which the gyrocentre of the particle is at rest. From a quantum mechanical viewpoint, (10.34) arises because momentum parallel to the magnetic eld and energy must be conserved (see Qu 2.). In the notation used here the resonance becomes (1 cos cos ) se / = 0. (6.10)

The properties of emission by relativistic particles discussed above imply that the emission occurs only on the cone dened by (6.4). Then (6.10) implies = se / sin2 . (6.11)

We now need to estimate the range of or s over which synchrotron emission predominantly occurs. We consider the temporal distribution of the radiation received by an observer. First note that because of the highly anisotropic angular distribution of synchrotron emission, signicant emission from a particle with pitch angle is seen by an observer only in the narrow range of angles about = . One pulse of radiation is received each revolution of the particle, with this pulse corresponding to the range of angles of order 1 during which the emission cone of the particle sweeps across the observer, as illustrated in Figures 6.2 and 6.3. Consider the simplest case = /2, which corresponds to the particle moving in a circle. The particle performs one gyration in a time 2/e . The time interval in which it is traveling toward the observer is a fraction of order 1 of this, that is, a time interval 2/e . Consider the pulse received when a particle with speed c moves directly toward the observer for a time t. Suppose that the particle is traveling in the x direction and that it starts radiating at t = 0 when it is at x = 0 and stops radiating at time t = t when it is at x = ct. (This starting and stopping of emission in the direction of the observer simulates the sweeping of the beam illustrated in across the direction to the observer.) The length of the wave train received by the observer is then ct ct ct/2 2 . That is, the length of the wave train is much shorter in the observers frame compared to the rest frame because in the observers the radiation from the particle emitted at t = 0 nearly catches up with the radiation emitted by the particle at t = t. The time taken for this wave train to move past an observer is t/2 2 . Now, since the time the particle spends travelling toward the observer each orbit is t = 2/e , the observed duration of the pulse is /e 2 . A pulse of duration t contains Fourier components < 1/t, so one expects synchrotron radiation to contain frequencies < 2 e . 74 (6.12)

The estimate (6.12) applies only to emission by particles with pitch angle = /2. However, it may be used to derive the corresponding result for arbitrary pitch angle by applying a Lorentz transformation. Consider a frame K in which a particle has pitch angle and Lorentz factor , and another frame K in which the particle has pitch angle = /2 and Lorentz factor . The frame K is moving relative to the frame K along the direction of the magnetic eld with velocity equal to the parallel velocity of the particle, that is, at c cos . The Lorentz factor of the transformation between the two frames is therefore (1 2 cos2 )1/2 1/ sin . Hence we have / sin , / sin . (6.13)
2 Then (6.12), now written < e / , implies

<

2 e sin .

(6.14)

It follows from (6.10) that the important harmonic numbers s involved in emission at are Thus, for ( sin )3 1, the emission is dominated by high harmonics and one is justied in treating s as a continuous variable. The discrete nature of the emission harmonics only becomes important when considering emission due to mildly relativistic particles (see 6.7).
1 3 s< ( sin ) .

(6.15)

6.3

The Emissivity for Synchrotron Emission

A detailed treatment of synchrotron emission may be developed as follows. The initial objective is to evaluate the emissivity in synchrotron emission. The emissivity is the power emitted per unit frequency and per unit solid angle, and it contains all relevant information on the distribution in frequency and angle. In view of (6.4), the angular distribution is singular to lowest order in 1/ , and may be described by a -function distribution, ( ). Specically, given the total radiated power, the power radiated per unit solid angle is found by multiplying by (cos cos )/2 . The power radiated per unit frequency can be determined from the power radiated at the sth harmonic. The total power may be written as a sum over the power at each harmonic.Very high harmonics are involved, and so s may be regarded as a continuous variable. The sum over s can then be replaced by an integral over s. This integral can be rewritten as an integral over by appealing to (6.11). One then identies the integrand of this integral over as the required power per unit frequency. Finally one needs to evaluate the power radiated at the sth harmonic and make the ultrarelativistic approximation to it. After a rather lengthy calculation, one can evaluate the power radiated at the sth harmonic exactly. It involves 75

Bessel functions, and these are approximated in terms of Airy integrals. This calculation may be performed separately for two linear polarizations, denoted here by and , corresponding to the direction of the electric vector in the radiation relative to the projection of the magnetic eld on the plane normal to the direction of propagation. (There is no circular polarization to lowest order in the expansion in 1/ .) The resulting expression for the emissivity is 3 q 2 e , (, ) = ( ) F , (, ), (6.16) 2 8 c 40 F
,

(, ) =

/c

dt K5/3 (t) (/c )K2/3 (/c ), (6.17)

2 c = 3 2 e sin .

The functions K (x) are modied Bessel functions, also called Macdonald functions.

0.8 0.6 0.4 0.2

0.5

1.5

2.5

Figure 6.4: The function F (R) dened by (6.19) The power emitted per unit frequency, found from (6.17) by integrating over solid angle and summing over the two states of polarization, is 2 3 q e sin F (/c), (6.18) P ( ) = 2c 40 with
1 F (R) = 2 [F (R) + F (R)] = R

dt K5/3 (t).
R

(6.19)

Expansions of the function F (R) for small and large arguments are given by 1/3 2/3 R 4 R 1 + , 1 2 (1/3) 2 3(1/3) 2 F (R) = (6.20) 1/2 R 55 R 1+ e + , 2 72R 76

for R 1, R 1, respectively. In between these limiting cases, there is a maximum at F (0.29) = 0.92. The function F (R) is plotted in Figure 6.4. A simple analytic approximation to it is F (R) 1.8R0.3 eR .

6.4

Emission by a Power-Law Distribution

Synchrotron emission is the emission mechanism for most radioastronomical sources, including supernova remnants, quasars and radio galaxies. In these sources the frequency spectrum is typically a power of the form I ( ) , where is the spectral index. This type of spectrum is interpreted in terms of emission from relativistic electrons with a power-law energy distribution. The emission due to a distribution f (p, ) = f (p)() of electrons is found by integrating the emissivity (6.17) over the distribution of electrons.This gives the emissivity per unit volume, J (, ). For a power-law distribution in energy it is convenient to introduce the energy spectrum N (), with = pc for relativistic particles, by writing

4
0

dp p2 f (p) =
0

d N (),

N () =

4 2 f (/c). c3

(6.21)

The energy spectrum is assumed to be a power-law of the form N () = Ka 0 for 1 2 , otherwise. (6.22)

In integrating the emissivity over this energy spectrum, the limits 1 , 2 can often be ignored, but at least one limit is essential in normalizing the number of particles, specically, one must have 1 > 0 for a > 1. The volume emissivity follows by integrating (6.17) over the distribution of particles 3 q 2 e sin () d N () F , (/c ) . (6.23) J , (, ) = 2 8 c 40 0 The integral over is performed as follows: set 1 = 0, 2 = ; note that the energy dependence of F (/c ), as given by (6.19) with (6.20), involves /c 2 and use this to change the variable of integration to the argument of the Macdonald functions; nally appeal to the standard integral
0

dx x K (ax) = 21 a1

1++ 2

1+ 2

(6.24)

with (1 + x) = x(x), The result is J


,

(1 x)(x) =

. sin(x)
(a1)/2

(6.25)

(, ) = A() j

(a) 77

2 3e sin

(6.26)

with K (mc2 )a+1 3q 2 e sin () , A() = 16 2 c 40 2/3 (a3)/2 3a 1 3a + 7 j (a) = , 2 a+1 12 12 a + 5/3 j (a) = j (a). 2/3

(6.27)

It follows that the spectral index a of the particle energy spectrum, and the spectral index of the synchrotron emission are related by = 1 2 (a 1), a = 2 + 1. (6.28)

The degree of linear polarization of the synchrotron emission is given by the dierence between the power in the two polarizations divided by their sum. Choosing to write the degree of polarization as positive if the component dominates, (6.26) implies rl = +1 a+1 = . a + 7/3 + 5/3 (6.29)

The relatively high degree of polarization is a characteristic signature of synchrotron emission. For a value of a = 3, which is typical for many sources, (6.29) implies rl = 3/4, that is, a degree of linear polarization of 75% in the direction orthogonal to the projection of the magnetic eld B in the source on the plane of the sky. This prediction is based on the seemingly unrealistic assumption that the magnetic eld in the source is uniform. In practice one expects the orientation of B to vary along the line of sight and across the source, and this would reduce the degree of polarization observed. The degree of polarization would be zero for a source in which the orientation of the magnetic eld is random. Nevertheless many synchrotron sources show relatively high degrees of polarization, for example, rl > 30% and a few have rl 70%.

6.5

Synchrotron Self-Absorption

The absorption process corresponding to synchrotron emission is called synchrotron self-absorption, or more simply, synchrotron absorption. The absorption coecient per unit length for the two states of linear polarization is
,

(, ) =

(2 )3 c2 2 2

d cos ()
1 0

d 2

(, )

d N () . (6.30) d 2

The absorption coecient (6.30) cannot be negative. This may be shown by partially integrating in (6.30) and using (6.17) and the properties of the Macdonald functions to show [2 , (, )]/ > 0.

78

For the power-law distribution (6.22), with the limits 1 = 0 and 2 = , (6.30) gives (2 )3 c K (mc2 )a 3q 2 e sin () , (, ) = 2 16 2 40 a/2 2 , (6.31) (a + 2)j , (a + 1) 3e sin with j , (a) given by (6.27). The transfer equation for synchrotron radiation including the eects of the polarization is written in the Mueller calculus, that is, in terms of the Stokes parameters I , Q, U , V . For simplicity the circularly polarized component is ignored (it is of order 1 compared with the linearly polarized components), and the natural modes of the medium are assumed to be circularly polarized. The transfer equation, including spontaneous emission, is then of the form I I Q 0 I d I Q = Q + Q I V Q , (6.32) ds U 0 U 0 V I where V = k describes the eect of Faraday rotation, with k the dierence in wave number between the two magnetoionic modes. The emission coecients are given by I = 1 J (, ) + J (, ) , c Q = 1 J (, ) J (, ) , c

(6.33)

and the absorption coecients by I =


1 2

(, ) + (, ) ,

Q =

1 2

(, ) (, ) .

(6.34)

Suppose that Faraday rotation is neglected in (6.32) so that only emission and absorption are included. Integration of (6.32) then gives I Q= I Q I Q 1 e(I Q )L , U = 0. (6.35)

The solution of (6.35) for I is plotted as a function of /c for given L and given N () in Figure 6.5. The intensity has a maximum as a function of frequency. The optically thin region, where absorption is unimportant, corresponds to frequencies above the maximum, where one has I (a1)/2 , as in (6.26). The optically thick region, where absorption is important, corresponds to frequencies below the maximum, where one has I 5/2 . The spectrum for an optically thick source may be interpreted in terms of a thermal-like spectrum I 2 with the temperature replaced by the energy of the particle, and with this energy rewritten as 1/2 which follows from c , cf. (6.20) and Figure 6.4. The degree of linear polarization rl = Q/I reduces to a simple form for (I Q )L 1 and for (I Q )L 1. The former limit corresponds to an 79

optically thin source with degree of polarization given by (16.24). The latter case corresponds to an optically thick source with the degree of polarization then given by rl = Q 3 = , I 6a + 13 for I L 1. (6.36)

The polarization implied by (6.36) for a = 3 is about 10%. Comparison of (6.29) and (6.36) shows that as a source becomes optically thick the plane of polarization ips through 90 and decreases in magnitude. Specic applications of opacity and polarization eects in synchrotron sources are addressed further in the next chapter.

log I
self-absorption turnover

2.5
optically thick

(a-1)/2
optically thin

log
Figure 6.5: The synchrotron spectrum of a source with electron distribution with a power law N () a .

6.6

The Razin Eect

The Razin eect refers to the suppression of emission at a frequency < p that occurs in a plasma with plasma frequency p . This eect causes a turnover in the spectrum that is somewhat similar to synchrotron absorption. However, the causes of the two types of turnover are quite dierent. The Razin eect is a true suppression and not an absorption eect. Electrons emit less eectively in a medium with refractive index less than unity than they do in vacuo. The 2 idea is that emission in a plasma with refractive index n = (1 p / 2 )1/2 is suppressed relative to emission in vacuo as the refractive index approaches zero, ie. when < p . Consider the Lorentz transformation between the frames K and K0 introduced in 2.1, cf. Figure 6.1. Recall that 2 k 2 c2 is an invariant under Lorentz 80

2 2 2 2 2 transformations, and is equal to p . In K0 one has 0 k0 c = p , so that 2 2 1/2 the refractive index in K0 is n0 = (1 p /0 ) . It follows that emission at 0 < p in K0 cannot occur. Consequently, the corresponding emission in K also cannot occur. The rst of (6.2) with n0 = 0 implies that the frequency 0 = p in K0 corresponds to = p in K . This leads one to expect that emission at < p is not possible in K , and while this is not strictly correct, emission at < p is strongly suppressed. A semiquantitative treatment treatment of the Razin eect involves introducing the eective Lorentz factor, as dened above, in (6.20) by replacing R by 3 R = 2s/3e sin3 = /c, 2 e = (1 + 2 p / 2 )1/2 ,

(6.37)

2 1/2 where the latter formula results from expanding e = 1/(1 e ) for 2 1, 2 2 /p 1. The presence of the plasma is unimportant for p when one has e . An important point in the following argument is that the emissivity is peaked around R 0.3, that is, at 0.3c, and it decreases strongly ( eR ) with R 1. For p inspection of (6.37) shows that R 1/ 3 increases with decreasing frequency. Thus the eect of the plasma is to cause R to have a maximum value as a function of at p . Provided that this maximum occurs at c the plasma has little eect on the synchrotron emission. However, if the maximum occurs at c (with c determined by ignoring the plasma), then the emission is conned to the exponentially weak region. The frequency at which this suppression eect sets in is found by solving /c = 1 for the value of at which the turnover occurs, and substituting this into = p . This gives the Razin Tsytovich frequency 2 RT = 2p /3e sin .

(6.38)

Below this turnover the spectrum rises must more steeply than for the selfabsorbed case illustrated in Figure 6.5. The Razin eect is of interest from a formal viewpoint. There is no known synchrotron source for which the eect is recognized as important.

6.7

Cyclotron emission

Cyclotron emission refers to gyromagnetic radiation from non-relativistic particles in the non-relativistic limit. The power radiated by a single charge is given by the non-relativistic limit of (6.8), namely P = e2 2 v 2 2 e sin . 60 c3 (6.39)

The angular distribution of the radiation and polarization are those characteristic of circular motion. The angular radiation pattern varies as 1 + cos2 , where 81

is the angle with respect to the magnetic eld (ie. = 0 is along the magnetic eld). Unlike synchrotron radiation, cyclotron emission may be highly circularly polarized. The ratio of circular to linear polarization is cos (see Qu. 5 below). The circular polarization is positive (right-handed) for negative charges gyrating in a right hand sense about the magnetic eld lines. Gryosynchrotron emission, which is emission from mildly relativistic particles, is important in some astrophysical situations. However, its full treatment is beyond the scope of this course. Many of the approximations used to describe emission in the non-relativistic and ultra-relativistic limits discussed above are inapplicable.

6.8

Questions

1. (Rybicki and Lightman 4.3) Show that the transformation of acceleration is ax ay az = = = a x , 3 3 a y 2 2 a z 2 2

u yv c2 u zv c2

a x , 2 3 a x 23

(6.40) (6.41)

where 1+ vu x . c2 (6.42)

If K is the instantaneous rest frame of the particle, show that a a = = 3a a


2

(6.43) (6.44) (6.45)

where a and a are the components parallel and perpendicular to the direction of v respectively. 2. Show that the resonance condition (10.34) arises in a quantum mechanical treatment as a result of conservation of momentum and energy. The energy eigenvalues of a relativistic particle (ignoring spin) are hc2 )1/2 . = n (p ) = (m2 c4 + p2 c2 + 2n|q |B (6.46)

Perpendical momentum is not conserved during emission, but both momentum parallel to the eld lines and energy are both conserved. Show 82

that this implies p = p hk, h. = n (p ) (6.47)

Now write n = n s and assume the changes in p , and n are all small. Then show that the expansion 2 + ), s + 1D h = (1 D 2 s s = h D s +k v p p (6.48)

reproduces (10.34) to lowest order in h . 3. Show that synchrotron absorption in vacuo cannot be negative. Negative absorption is only possible for d(2 grate
, ,

)/d < 0. Partially inted N () (6.49) d 2

(, ) =

(2 )3 c2 2 3

d cos ()
1 0

d 2

(, )

and show that negative absorption is possible only for d(2 , /d < 0. Using the relations R = /c = 2s/3 3 sin3 , = /mc2 and the relations (10.40) and (10.41), use the following and K (R) > 0 to complete the proof: d(2
,

/d) = 2R

3 1 2 K5/3 (R) + 2 K1/3 (R) 2 3R K2/3 (R)

(6.50)

4. Show that the reversal of the sense of polarization as a synchrotron source becomes self-absorbed occurs at a frequency determined approximately by e
L

2 3a + 5

(6.51)

5. The polarization vector e of dipole emission is determined in terms of d( ) by noting that the energy emitted is proportional to |e d( )|2 and hence the polarization is along the projection of d( ) on the plane orthogonal to the direction of emission. Determine the polarization in the following cases as a function of the angle between and the z axis. a) The system is oscillating such that d( ) is real and along the z axis. b) The system is rotating about the z axis such that d( ) is proportional to (1, i, 0), where the signs refer to the right and left hand senses of polarization, respectively.

83

Chapter 7

Astrophysical Synchrotron Sources


Reference: Rybicki & Lightman, Chapter 6
Having derived the characteristics of synchrotron emission, we apply this knowledge to some astrophysical situations.

7.1

Summary

The shape of the frequency spectrum of a synchrotron source is determined by the energy distribution of the emitting particles and by the properties of synchrotron emission. In the simplest case, the frequency distribution is a declining power law over the range of interest, so the intensity I ( ) and this is interpreted as being due to a power law electron distribution, N () a . The power law indices are related by (6.26): = 1 (a 1), 2 a = 2 + 1. (7.1)

Astrophysical synchrotron sources with moderately steep spectra are usually quite well tted by power laws. A wide range of values of is found, with = 1 implying a = 3 being a value often chosen for illustrative purposes. It is pertinent to ask whether it is possible to invert the frequency spectrum to nd the particle energy spectrum, but there is no simple answer. Such an inversion procedure is subject to both physical and mathematical uncertainties. The physcial uncertainties include the possibility that the distibution for the emitting particles is anisotropic, i.e. is a function of both energy and pitch angle. It is certainly not possible to invert a synchrotron spectrum to obtain an anisotropic particle distribution. This diculty is discussed below in connection with the evolution of synchrotron spectra. The mathematical uncertainty concerns how well posed the inversion problem is. The synchrotron spectrum 84

may be regarded as a convolution of the energy spectrum of the particles and the synchrotron spectrum of an individual particle. When the frequency spectra of the individual particles is intrinsically narrow there is an approximate oneto-one relation between the particle energy and the frequency of emission, and so the inversion of the frequency spectrum to nd the energy spectrum is well dened. However, this is not the case if the frequency spectra of the individual particles are broad. The emission at a given frequency could then be dominated by emission near the mean frequency of particles of an appropriate energy, but it could also be due to the low frequency emission from higher energy particles, the high frequency emission from lower energy particles, or any combination of these. As noted following (6.20), the single-particle spectrum may be approximated by F (R) = 1.8R0.3 eR with R = C/2 , where C is a constant. In this approximation the integral for the frequency spectrum due to a particle spectrum N () is of the form

J ( ) =
0

C d F (R) N () = 2

F (R) dR 3/2 N R

C R

(7.2)

This has the form of a Laplace transform, and is, in principle, easily inverted to nd N (). However in practice the inversion problem is often not well posed because the requirement that the function is sampled over a wide frequency range is not met. Synchrotron spectra sometimes exhibit a low-frequency turnover, i.e. the frequency spectrum at low frequencies increases with increasing frequency, then turns over, and at high frequencies is decreases with increasing frequency. There are several possible causes of such turnovers, some of which are illustrated schematically in Figure 7.1. The Razin eect and external free-free absorption (collisional damping) cause exponential low-frequency cutos. Synchrotron self-absorption implies a low frequency spectrum 5/2 . Internal free-free absorption inplies a low-frequency spectrum 2 . Induced Thomson scattering (induced scattering of photons by non-relativistic electrons) implies a low-frequency spectrum 1/2 . A low-energy cuto in the electron energy spectrum leads to a low-frequency spectrum 1/3 , as implied by the spectrum for an individual particle (see 6.20). A practical problem in interpreting any synchrotron source is that the nergy of the emitting particles and the magnetic eld cannot be determined independently solely from a power-law synchrotron spectrum. This is because of the dependence B2 implied by (6.14). However, one can try to estimate the magnetic eld using some some spectral features (e.g. breaks in the spectrum due to synchrotron ageing) or using equipartition arguments.

7.2

The spectrum due to synchrotron radiation

A quantitative interpretation of synchrotron emission is required to interpret observations of specic sources. Here we assume the source to be a slab of 85

log I

low energy e energy cutoff

0.3

=(a-1)/2
synch self-abs'n

2.5

optically thin

induced Thomson scattering

1/2

internal free-free

exponential cutoff
(Razin effect and external free-free absorption

log
Figure 7.1: The various causes of spectral turnovers. A synchrotron spectrum is illustrated in the optically thin region at high frequency. thickness L along the line of sight, and that the particle distribution is isotropically distributed in pitch angle, i.e. () = 1 so f (p, ) = f (p). The specic intensity from an optically thin source is identied by writing I ( )d = LJ ( )d , with J ( ) the emissivity per unit volume given by the sum of the two polarized components (6.26). Then one nds (a1)/2 (a1)/2 LK (mc2 )1a 3e2 B sin 3 sin , (7.3) I ( ) = j (a) 4c 40 2 B with j (a) = j (a) + j (a), and where the cyclotron frequency is B = e = 2.8 1010 2 B 1T Hz = 2.8 106 B 1G Hz. (7.4)

In practice there is usually no information about , the orientation of the magnetic eld in the emission with respect to the line of sight, and it is appropriate to average over all directions. Using the result
+1

d cos (sin )(a+1)/2 = 1/2


1

( a+5 4 ) , ( a+7 4 )
(a1)/2

(7.5)

one nds I ( ) = A(a)LK (mc2 )1a where e2 B 40 c 2 3B , (7.6)

(a1)/2 a+5 +19 1 ( )( 3a12 )( 3a 32 12 ) . A(a) = 4 8 (a + 1)( a+7 4 ) 86

(7.7)

The quantity K (mc2 )1a has the dimensions of number density. The number density of radiating particles n and particle energy density Upar are given by K K a a a Upar = . (7.8) 1 2a 2 1 1 2 2 a1 a2 1 The simplest way to estimate the lower and upper energy cutos 1 and 2 respectively is to assume that all the synchrotron emission from the particles at these energies occurs at an angular frequency /c = 0.31 where c = 1.5e 2 sin . One then equates cutos in the observed emission spectrum with cutos in the energy spectrum. Now the ux density from a spherical source of radius Rs at a distance Ds is related to the intensity by n= F ( ) = I ( ) Rs Ds
2

2 I s , 4

(7.9)

where s = 2Rs /Ds is the angular size of the source. So, if we know the ux density of a source at a given frequency, together with its angular size, then we can calculate the intensity, so we know the left hand side of (7.6). If we also have some information regarding the depth L we can use (7.6) to nd K . Numerically, one has e2 /40 c = 7.7 1037 W Hz2 , L is often expressed in parsecs where 1 pc= 3.09 1016 m, and F is usually measured in janskys, 1 Jy= 1026 W m2 Hz1 . The magnetic eld and low frequency turnovers If there is a low-frequency turnover in the spectrum an estimate of B can be obtained by assuming that the turnover is due to synchrotron self-absorption. The brightness temperature of the source at the frequency m corresponding to the maximum ux Fm is kTB (m ) = c2 2 c2 2 = . I Fm s m 2 2 2m m (7.10)

For a self-absorbed source this is roughly equal to the energy of the particles radiating at m which is m mc2 2mm 0.47 eB sin
1/2

(7.11)

Equating (7.10) and (7.11) then gives 3 m3 5 2 4 (7.12) F . 0.94 e m m s The sin term is generally ignored because of the inherent uncertainties involved in the overall procedure and since we usually have no way of estimating . This estimate is extremely sensitive to the frequency at which the peak of the spectrum occurs, and the angular size of the source. Estimates of the magnetic eld made in this way are therefore subject to large uncertainties. B sin 87

Equipartition In the absence of a low-frequency turnover, and of any other information, an argument based on the equipartition of energy between the radiating particles and the magnetic eld is often used to estimate B . One writes B2 , (7.13) 8 and assumes = 1 or some other specic value. There is a weak theoretical argument to support (7.13), namely that the total energy density Upar + B 2 /8 for the observed synchrotron emission should be a minimum. From (7.6) one nds K B (a+1)/2 for a xed synchrotron spectrum, and then expressing the particle energy density in terms of B implies Upar B 3/2 . Minimizing Upar + B 2 /20 then implies 3Upar /2 + 2Umag = 0, which is simply (7.13) with = 4/3. The observational evidence for (7.13) is also weak. For example, in the interstellar medium the total energy density in cosmic rays is in approximate equipartition with the interstellar magnetic eld, but the energy density in cosmic ray electrons is such that = 0.03. Upar = Umag Umag =

7.3
7.3.1

Simple source models


Slab geometry

Here we consider aspects of the polarization in synchrotron sources. In a slab model, one integrates (6.32) through a source of thickness L in general. Consider the simpler cases in which only one of the last two terms in (6.17) is retained. First, suppose that the absorption term is neglected. Then integration of (6.32) gives Q Q I = I L, Q = sin(V L), U = [1 cos(V L)]. (16.30) V V This result describes the eect of spontaneous emission in a medium in which signicant Faraday rotation occurs. It follows from (16.30) that in the limit of large Faraday rotation, that is, for V L 1, Q and U do not exceed Q /V . This implies that the degree of polarization is of order Q /I V L, which is a factor 1/V L smaller than in the absence of Faraday rotation. This reduction in the degree of polarization due to Faraday rotation is called Faraday depolarization. The actual spectrum of a source which has a low-frequency turnover due to self-absorption generally diers from the idealised slab model spectrum shown in Figure 7.2. Observed spectra from unresolved sources have broader peaks than implied by the slab model because dierent parts of the source become optically thick at dierent frequencies. This is illustrated schematically in Figure 7.3. A real source which is optically thick remains optically thin near the limb because the line of sight through the source is shorter there, so the emission below the peak in the spectrum is always greater for a real source than for the idealised slab model. 88

7.3.2

Core-jet model

The radio emission from some AGN is interpreted in terms of a core-jet model, as illustrated in Figure 7.2. Suppose that the core is optically thick with a spectrum of relativistic particles of the form N ( ) = K 2 for min < < max . The jet is assumed to move with constant speed within a cone. Particle acceleration is assumed to take place continuously, so that min , max remain constant. Then conservation of particles implies K r2 . Assuming that the azimuthal magnetic eld dominates, it scales as B r1 .

optically thin

optically thick low high

Figure 7.2: A simple model of a nonrelativistic core-jet source. The core is the unresolved optically thick inner region, shown shaded. The synchrotron emission from an optically thin and an optically thick portion of the jet then scale as S (, r) KB 1.5 0.5 r3 (r)0.5 , S (, r) B 0.5 2.5 r2 (r)2.5( , 7.14)

respectively. Both relations (7.14) depend only on the combination r, rather than separately on and r. The maximum emission occurs approximately where the two estimates in (7.14) are equal. It follows that one has Smax ( ) = const, max r1 . (7.15)

The predicted spectrum is at ( = 0), as illustrated in Figure 7.3. This model is roughly in accord with the observations, except that it is nonrelativistic.

7.4

Evolution of Synchrotron Spectra

The spectrum of a synchrotron source can evolve due to synchrotron losses, adiabatic expansion or compression, and a variety of other processes. The eect of synchrotron losses may be considered in two separate cases: when a continuous 89

log S()

log
Figure 7.3: A at integrated spectrum, shown by the solid line, results from the model discussed in the text in which the emission from each point in the source has the self-absorbed shape indicated by the dashed curves. supply of radiating electrons is balanced by synchrotron losses, and when an initial electron spectrum evolves with time. The simple results are used. First, relativistic particles emit in the forward direction, so the momentum change is in p, and not in the pitch angle, . Second, the rate of energy loss must equal the power radiated. Hence, one nds = b() 2 , b() =
2 2 e 2 2 e sin . 5 3(40 )m2 ec

(7.16)

For the present we assume that the distribution is isotropic and average over pitch angle using sin2 = 2/3 to obtain b = b() = e4 B2. 5 90 m4 c e (7.17)

The evolution of a distribution of electrons subject to synchrotron losses is described by N (, t) = [N (, t)] + Q(, t) L N (, t), t (7.18)

where the three terms on the right hand side describe synchrotron losses, a source of electrons and escape of electrons, respectively. Equation (7.18) can be integrated explicitly after Laplace transforming. In the special case where the source term is a power law that is independent of time, Q() = Aa , the solution is N (, t) = AeL /b b2
/(1bt)

d eL /b a ,

(7.19)

90

with the upper limit replaced by for bt > 1. If electrons cannot escape, ie. when L = 0, equation (7.19) simplies to N (, t) = Aa (a 1)b 1 (1 bt)a1 for b1 t 1, (7.20)

for b1 t > 1.

This solution is illustrated in Figure 7.4. Synchrotron losses cause the power law spectrum a to steepen at high energies to (a+1) ; the bend occurs at an energy = 1/bt that decreases with increasing time. This bend in the energy spectrum implies a steepening of the synchrotron spectrum such that the power law index increases by 0.5 from lower to higher frequencies.

log N(,t)

-a -(a+1)

log

Figure 7.4: A plot of the solution (7.20) showing the increase in power law index from a to a + 1 at 1 = 1/bt. The synchrotron half life is the time in which the energy of an electron decreases by a factor of one half due to synchrotron losses. Solving (7.16) for at time t, given = 0 at t = 0, one nds 1 1 + bt. = 0 Hence the half lifetime is identied as t1/2 = 1 = 5 108 sin2 b B 1G
2

(7.21)

mc2

s.

(7.22)

The energy at which the spectral break occurs is that for which the synchrotron half life of the electrons is equal to the time elapsed. One specic model for a class of extragalactic synchrotron sources is based on evolution due to synchrotron losses following multiple injections of electrons. 91

The model is intended to explain synchrotron spectra with two bends, as illustrated in Figure 7.5. The bend at lower frequencies is attributed to the change = + 0.5 due to synchrotron losses on a timescale which is long compared with the interval between injections. Then the multiple injections may be treated approximately as a continuous injection, so that (7.20) applies. The higher frequency bend is = 4/3 + 1, and is due to an eect not included in the foregoing discussion. Over time the electron distribution becomes highly anisotropic because the synchrotron lifetime is inversely proportional to sin2 . Thus particles with large pitch angles lose energy more rapidly than particles with small pitch angles. Therefore, at a time long after the last injection, the highest energy particles will all have small pitch angles, , and so radiate only at small angles . The spectral index = (2 + 1)/3 is found by averaging the spectrum due to this highly anisotropic component over all directions, i.e. over , which corresponds to assuming that all orientations of the magnetic eld occur randomly throughout the source. A criticism of this model is that pitch angle scattering should be ecient and would prevent the highly anistropic distribution from developing.

log I()

'

''

Figure 7.5: The synchrotron spectrum for a model based on multiple injections. Synchrotron losses cause the bend from to = + 0.5 at 1 and the bend to = 4/3 + 1 at 2 is due to the particle distribution becoming highly anisotropic at high energies. Another important eect on the evolution of synchrotron sources is adiabatic expansion. The simplest model, called the synchrotron bubble, is one in which a source expands isotropically. In this case L1 , and for a source which expands at constant speed it follows that the energy, , of a particle at time t is related to its energy, 0 , at time t0 by = 0 (t0 /t). Then if there is no continuing supply of radiating electrons, the total number of electrons is a constant and integrating over an initial power law spectrum N (0 , t0 ) = K0 0 92

a implies K (t)1a V = K0 1 V0 where V is the source volume. The evolution 0 fo N (, t) follows trivially, and one nds

N (, t) = K (t)a = K0

t t0

(a+2)

a .

(7.23)

Substituting this form of K (t) into the synchrotron formula, with L t and B B L2 t2 , then implies I L2(a+1) . Thus the emission at xed frequency from a xed number of particles has a strong dependence on scale length L. This eect is of interest when a localized region of a source is subject to compression, so that the parameter L, which is characteristic of the dimension of the source, decreases. The synchrotron emission at xed frequency then increases as L2(a+1) , e.g. for a = 3 one nds I ( ) L8 .

7.4.1

Synchrotron lifetime in astrophysical sources

In some astrophysical synchrotron soures, the synchrotron half lifetimes of the radiating electrons are so long that it is quite possible that the electrons were rapidly accelerated and the source subsequently expands adiabatically (the synchrotron bubble model). In others, notably the Crab Nebula and certain astrophysical jets, the synchrotron half lifetimes are so short that they provide compelling evidence that the radiating electrons must be undergoing continuous acceleration. One reason is that the diusion of electrons away from an acceleration site is quite slow: the net ow speed is expected to be of order the Alfv en speed. Thus it is plausible that electrons are accelerated essentially at the point where they are observed to radiate, rather than being accelerated at some remote site and time and propagation to the point at which they are observed. The slow diusion of energetic particles away from an acceleration site leads one to expect the synchrotron spectrum to vary away from the site due the synchrotron and expansion losses.

7.5

Questions

1. Show that the average rate of energy loss by an electron in an isotropic distribution (ie. averaged over pitch angle) is = b2 , b= e 2 2 e . 5 90 m2 ec (7.24)

The evolution of the energy spectrum N (, t) for electrons in a synchrotron source is modelled by an equation of the form N (, t) = [N (, t)] e N (, t) + Q(, t), (7.25) t where Q(, t) is a source term and e N (, t) represents a loss term due to escape of electrons from the source. Solve (10.44) in the stationary case Q() . 93

94

Chapter 8

Compton Scattering
Reference: Rybicki & Lightman, Chapter 7
In this chapter we investigate the scattering of radiation by particles, usually electrons. The classical theory of this scattering is known as Thomson scattering, and its generalisation to include quantum and relativistic eects is referred to as Compton scattering. In both these processes there is a redistribution of the angular pattern of the radiation, and the radiation imparts energy to the electron. This results in a decrease in the energy of the scattered radiation, although this decrease is negligible under many cases of practical interest. The scattering of radiation o relativistic particles is known as inverse Compton scattering; in this process the particles usually impart energy to the radiation. In this Chapter we investigate both Compton and inverse Compton scattering processes. We consider their application to astrophysical sources by considering the limits in the brightness temperature of a source due to inverse Compton scattering. Implications of the synchrotron self-Compton model are also explored.

8.1

Thomson Scattering

Here we review classical Thomson scattering, which is the scattering of one electromagnetic wave into another electromagnetic wave by a charged particle. The electron and photon do not exchange energy, but there is a change in the direction when the electron re-radiates the electromagnetic energy of the photon. Unbound electrons are hard to hit, and their scattering cross section is small relative to that of a hydgrogen atom. Thus photoionisation usually dominates over Thomson scattering, except when the medium is hot, and is already ionized. Thomson scattering applies when the wavelength of the indicent radiation is much smaller than the size of the atom or relevant particle. Scattering in the opposite regime, where the wavelength exceeds the size of the atom, is known as Rayleigh scattering. The strength of Rayleigh scattering scales as 4 . Rayleigh

95

scattering is responsible for the colour of the blue sky during the day and its red colour during sunsets. It is also responsible for its linear polarization. Thomson scattering is viewed as a process in which the charged particle is accelerated by an electromagnetic eld incident on it and subsequently emits radiation as a result of the acceleration. The present treatment assumes that the energy of the emitted photons h is less than mc2 . It also assumes that the particles are moving non-relativistically, so that the emission may be treated as electric dipole emission using the Larmor formula. Suppose radiation in the form of a plane wave, E(t, x) = E0 cos(t k0 x), (8.1)

is incident upon a charge, q . The charge oscillates as a result of the radiation. In the non-relativistic approximation the magnetic force on the particle is negligible, hence the acceleration is a = q E(t, x)/m. Subsitution into (3.22) gives the time-average power : P = q 4 |E0 |2 . 120 m2 c3 (8.2)

The angular distribution of the power is given by inspecting the Poynting vector (3.17). For radiation scattered by an angle from in the plane dened by the the wavevector and the direction of the electric oscillations, E0 /E0 , the power per solid angle is 3 q 4 |E0 |2 dP cos2 . = P cos2 = d 8 32 2 0 m2 c3 (8.3)

Now, since the incident ux is S = c/8 |E0 |2 , it is convenient to express the power scattered by solid angle in terms of the dierential cross section between incident and outgoing angles d/d: d dP = S . d d (8.4)

Using (8.3) we obtain an explicit expression for the dierential scattering of angles d into d: d 2 = r0 cos2 , d (8.5)

where r0 = e2 /40 me c2 2.82 1015 m is the classical electron radius. Both the scattered power and the dierential cross section are independent of frequency under the assumptions made here. The total cross section is found by integrating over all solid angles d: T = = d d = 2 d 8 2 r . 3 0 96
1 1 2 d sin r0 sin2

(8.6)

This is known as the Thomson cross section. We derive the polarization properties of the scattered radiation due to unpolarized incident radiation. Consider radiation with wavevector along the z -axis. One can consider unpolarized radiation as a sum of polarized contributions, with the magnitude of the electric eld in the x and y directions being equal, and with the phase between the oscillations in the two directions a random variable. Now the power scattered from the radiation polarized in the x-direction is q 4 |E0 x |2 dPx cos2 x , = d 32 2 0 m2 c3 and the power scattered from radiation polarized in the y -direction is q 4 |E0 y |2 dPy cos2 y , = d 32 2 0 m2 c3 (8.8) (8.7)

and y is the y component. Now we take the where x is the x component of n special case y = 0. Setting E0 x = E0 y for unpolarized radiation and summing the contributions from both polarizations we have r2 dP = 0 (1 + cos2 ). d 2 (8.9)

The dierence of the two polarizations gives us the degree of linear polarization for the scattered radiation from unpolarized incident radiation: = 1 cos2 I1 I2 = . I1 + I2 1 + cos2 (8.10)

We note that both (8.9) could have been inferred directly from (3.20).

Figure 8.1: The electrons motion is conned to a plane normal to the direction of the incident radiation. To an observer looking directly along the direction of a beam on a particle, all directions on the sky look identical by symmetry, and there is no nett polarization. However, the radiation is fully polarized when looking perpendicular to the incident wave.

8.1.1

Compton Scattering

Quantum mechanical considerations alter the classical derivation of Thomson scattering outlined above in two ways: 97

Thomson scattering assumes elastic collisions. Under certain circumstances the electron recoil cannot be neglected. The scattering cross section is reduced as the energy of the incident photon, h , becomes comparable to or exceeds the electron rest mass energy, mc2 . We now consider each of these eects in detail. Conservation of four-momentum requires that the scattering obey Pei + Pi = Pef + Pf . (8.11)

Assume the electron is initially at rest and taking the directions of the incident and scattered radiation to be ni and nf respectively, we have Pei Pef Pi Pf = = = = (mc, 0, 0, 0), (E/c, px , py , pz ), i (1, ni )/c, f (1, nf )/c. (8.12)

Solving these equations according to (8.11) yields the energy of the scattered radiation i f = , (8.13) i 1 + mc2 (1 cos ) where is the angle between the wavevectors of the incident and scattered wavefronts. There is a critical energy or, equivalently, wavelength at which the inelasticity of the scattering becomes important. This is seen more clearly by rewriting (8.13) as f i = c (1 cos ), where c = h = 0.024 A for electrons, mc (8.15) (8.14)

is the Compton wavelength. The scattering is nearly elastic for wavlengths longer than the Compton wavelength. Equation (8.14) shows that the Compton scattering always results in a loss of energy to the radiation. Since 0 < < , the wavelength of the scattered radiation is always larger than that of the incident radiation. Quantum eects also aect the cross section of the scattering process. The scattering cross section is reduced if the energy of the outgoing photon, 1 is comparable to the energy of the incident photon, . The dierential cross section is then given by the Klein-Nishina formula r 2 1 d = 0 d 2 1 + sin2 . 1 98 (8.16)

In the non-relativistic and ultra-relativistic regimes the eective cross section is approximated by = T 1 2x + x
1 26x2 5

3 8

ln 2x +

1 2

for x 1, for x 1

(8.17)

where x = h/mc2 . The Klein-Nishina formula thus reduces to the Thomson cross section in the classical limit, but it is much smaller than this in the ultrarelativistic regime. In the foregoing section we considered scattering of radiation o an electron initially at rest. However, in general, the electron will possess some kinetic energy and if this is high enough, it may impart of some of its energy to the radiation. This is the subject of the following section.

8.2

Inverse Compton Scattering

Inverse Compton scattering refers to the scattering of radiation by relativistic particles, as depicted in Figure 8.2. In this process it is possible for photons to gain energy by scattering o electrons.

Figure 8.2: Inverse Compton scattering of low energy photons o energetic electrons upscatters the photons to high frequencies. The scattering is considered by transforming from the laboratory frame, K , where the particle is moving with Lorentz factor , to the rest frame of the electron, K0 . In the observers frame K the incident radiation makes an angle i to the direction of the moving electron, and the scattered radiation makes 99

x0

K0 - particle rest frame

K - laboratory frame

i ' e-

f'
f' e-

i f

i'
z0

i
z

Figure 8.3: The angles and energies of the incoming and outgoing photons in the rest frame of the particle and in the laboratory frame.
an angle f . In the electron rest frame these angles are i and f respectively (see Figure 8.3). The azimuthal angles of the incident and scattered radiation in the rest frame are denoted i and f respectively. Lorentz transformation of the electron energy shows that the apparent energy of the radiation in the rest frame is

i = i (1 cos i ),

(8.18)

and the energy of the scattered radiation transformed back into the observers frame is
f = f (1 + cos f ).

(8.19)

Equation (8.13) provides the relationship between energies in the rest frame of 2 the particle, which, expanded for small i < mc , becomes i (1 cos ) , mc2 cos = cos f cos i + sin f sin i cos( f i ).
f = i 1

(8.20) (8.21)

Equations (8.18) and (8.19) imply that the ratios of the incident energy in the observers frame, in the rest frame before scattering, and in the observers frame after scattering are roughly 1 : : 2. (8.22) Thus a particle can increase its energy by a factor 2 in a single encounter. Furthermore, since the encounter in the electron rest frame depends only on one power of , the scattering can occur in the Thomson regime, where the scattering cross section is large compared to the Klein-Nishina regime applicable at higher energies. 100

8.2.1

Average scattered power

The average power scattered by an electron requires an average over the range of angles of the incoming and outgoing photons. Let us dene a photon occupation number N (k), such that the number of wave quanta with wavenumbers between k and k + dk in a box of size d3 x is N (k)d3 xd3 k/(2 )3 . The energy volume density in this range of momenta hk, is then W (k) = h N (k)d3 k/(2 )3 . Now the photon occupation number is a Lorentz invariant, so one has N (k)d3 k = N (k )d3 k . (8.23)

The energy loss due to inverse Compton scattering in the rest frame of the electron is dE = cT dt ( h )2 N (k ) d3 k (8.24)

Using (8.18) and the fact that dE /dt = dE/dt, the rate of energy loss is dE dt = = where Uph = ( h )2 N (k) d3 k, (8.26) cT 2 (1 cos )2 cT 2 (1 + 2 /3)Uph , ( h )2 N (k) d3 k (8.25)

is the initial mean energy density in radiation. Since 2 1 = 2 2 the inverse Compton scattered power is PCompt = 4 T c 2 2 Uph . 3 (8.27)

One can compute that total power scattered by a distribution of relativistic electrons by integrating (10.25) over the distribution of electron energies.

8.3

Synchrotron Self-Comptonised Emission

A specic but important model is when the electrons responsible for synchrotron emission in an astrophysical source can also inverse Compton scatter some of this radiation to higher energies. The other case is external Compton scattering, when high energy particles scatter externally incident radiation, and is not discussed here. The power radiated in synchrotron emission may be written in the form (c.f. (6.8) averaged over pitch angle) Psynch = 4 T c 2 2 Umag , 3 101 (8.28)

where Umag = B 2 /20 is the magnetic energy density. This has the same form as (10.25). Thus the ratio of power due to synchrotron emission to that by inverse Compton scattering is given by the ratio of the magnetic energy density to the photon energy density Umag Psynch = . PCompt Uph The power radiated per electron is Ptot = 4 T c 2 (Umag + Uph ), 3 (8.30) (8.29)

with two characteristic emission frequencies corresponding to the two radiation processes: = 0.5e 2 sin 4 2 3 for synchrotron emission with h = Urad /nph , (8.31)

where nph is the number density of photons. The loss of energy due to these two processes leads to the electron cooling on a timescale tcool = 3 m e c2 me c2 = . Ptot 4T c (Umag + Urad ) (8.32)

Note that in describing the energy losses in (8.30) we have not explicitly included corrections to the radiation energy density Urad due to the contribution of synchrotron photons that have been upscattered by the inverse Compton process. This eect of this correction is discussed below.

8.3.1

The Inverse Compton Limit for Synchrotron Sources

The brightness temperature of a synchrotron source is limited by inverse Compton scattering. This limitation occurs because synchrotron seed photons can be scattered multiple times. This causes a large increase in the energy density of the radiation eld, which eventually leads to the inverse Compton catastrophe. Consider the power in the radiation eld, Urad . It consists of the energy density due to photons emitted by the sycnchrotron radiation. However, it also consists of the energy density due to photons that have been inverse Compton scattered once, the photons that have been inverse Compton scattered twice, and so on. Thus the total power emitted can be written as a sum over all generations of scattered photons P =
2 4 Urad Urad + 2 T c 2 Umag 1 + + . 3 Umag Umag

(8.33)

The total power lost by a single electron is geometric series which is proportional to 1/(1 Urad /Umag ). The energy losses due to inverse Compton scattering are asymptotically large as Umag /Urad approaches unity. This is known as the 102

inverse Compton catastrophe. It implies that the radiation energy in a source cannot exceed the magnetic energy density. If this were to happen radiation losses in the source would become so large that the source would very eciently (ie. quickly) radiate away the extra energy until the radiation energy density is again suciently low. The inverse Compton limit is commonly expressed as a brightness temperature limit. The energy densities in the two elds are Umag = B2 20 Urad 4 3 kTB . c3 (8.34)

The brightness temperature of an electron of energy me c2 is simply TB me c2 /k . From (8.31) the frequency of emission is e 2 so one has B kTB m e c2
2

(8.35)

Equating the energies in the two elds (8.34) yields TB = = me c15/7 k (8r0 B )1/7 1.0 1012 B 104 T
1/7

(8.36) K. (8.37)

The brightness temperature of the inverse Compton limit is remarkably robust because it is highly insensitive to the model parameters. Modications, such as including the proper Klein-Nishina cross-sections for scattering by relativistic particles, do little to alter the actual value of the inverse Compton limit. The optimal emission frequency is also determined by the model using (8.35), and yields opt = B c 8r0 B
2/7

= 1.2 1011

B 104 T

5/7

Hz.

(8.38)

One has TB 1/2 (I 5/2 ) in the optically thick regime ( < opt ) and I in the optically thin regime ( > opt ). There is no direct evidence that this process actually limits the brightness temperature of astrophysical sources in that no catastrophe has ever been seen, but there is strong indirect evidence in that relatively few synchrotron sources 12 with TB > 10 K are observed (for exceptions see the following chapter on interstellar scintillation). However, it has recently been suggested that energy equipartition may limit the intrinsic brightness temperature of the synchrotron 12 sources to TB < 10 K, but no detailed mechanism has yet been proposed.

8.4

Eectiveness of Compton Scattering

It is useful to quantify the importance of Compton scattering in a given medium. The value of the Compton Y parameter indicates to what extent the energy of 103

a photon is changed by Compton scattering, and is dened as Y Nscat , (8.39)

where is the mean fractional change in energy per encounter and Nscat is the typical number of times the photon is scattered before escaping the medium. For Y 1 Compton scattering does not signicantly alter the photon energy, while for Y > 1 Compton scattering appreciably alters the energy of a photon. We estimate both of the factors that contribute to Y below. For a thermal distribution of electrons the fractional change in energy is computed from (8.20) which, averaged over angle, yields
f i i = = . mc2 i

(8.40)

Now, we want to know the change in energy in the lab frame. We write the fractional change in energy in a similar form as (8.40) but with a correction term: = 2 + , mc (8.41)

where the factor is to be determined. This is done by assuming that the photons and electrons in the scattering medium are in equilibrium. The photons have a Bose-Einstein distribution rather than a Plank distribution because the photons can neither be created nor destroyed by the scattering. In the non-degenerate limit, the appropriate photon energy distribution is the ultrarelativisitic thermal distribution N () 2 exp /kT , with = N () d N () d 2 = 3 kT = 12k 2 T 2 . (8.43) (8.44) (8.42)

The equilibrium requirement that no nett energy is transferred between the electrons and photons implies = 0, and allows us to nd by taking the average of (8.41) = 0 = 2 + . mc2 (8.45)

Relations (8.44) then imply that the mean energy imparted to the photon per encounter is (4kT ). (8.46) = mc2 104

Thus energy is on average imparted to the photons if the electron temperature T exceeds /4k via inverse Compton scattering. On the other hand, electron temperatures lower than this cause energy to be transferred from the photons to the electrons via Compton scattering. The nett transfer of energy when the electrons are highly relativistic is easier to compute. Combining relations (8.18) and (8.19) and averaging over angles we have f i 2 (1 cos )2

(8.47)

hence for 1 4 2 . 3 (8.48)

If the ultrarelativistic electrons are thermally distributed we can assign them a temperature and 2 = 2 = 12k 2 T 2 . Hence the average energy imparted to the photon in this case is 16 kT m e c2
2

(8.49)

We now estimate the number of scattering events encountered a photon as it escapes the scattering medium. In an optically thick medium a photon is scattered many times before it exits the scattering region. The number of scatterings is determined from the mean free path of the photon, lmf p and by considering the photons trajectory as a random walk. Consider a photon initially at the origin. Its position after N scatterings is the sum of discrete jumps from one scattering event to another xN = xN 1 + xN . (8.50)

Clearly, for isotropic scattering one has xN = 0 hence xN = 0 (ie. the mean position is its initial position). However, a succession of scatterings can cause the photon to leave the scattering medium. It is therefore useful to determine the particles root-mean-square di 1/2 stance from the origin. If each jump has a root mean square length x2 = N lmf p , then it is easily shown that the root mean square displacement from the origin after N jumps is 1/2 (8.51) x2 = lmf p N. N Now, in a medium of thickness L the optical depth is roughly L/lmf p . Equation (8.51) implies that the particle undergoes N (L/lmf p )2 scatterings before escaping the scattering region. So, for an optically thick medium in which the particle is scattered many times before exiting, the number of scatterings is N 2. 105 (8.52)

In the opposite limit, when the optical depth of the medium is less than unity, most particles escape without being scattered. The mean number of scatterings per photon is of order the fraction of photons that undergo a single scattering 1 e 1. To summarise, in a medium of arbitrary optical depth the mean number of scatterings is N 2 < 1 . 1 (8.53)

The actual optical depth of the medium is related to its size L, density, and the electron scattering opacity, scat : scat L. For ionized Hydrogen the scattering opacity is scat = T = 0.04m2 kg1 . mp (8.55) (8.54)

8.5

What is the energy spectrum of photons scattered o relativistic electrons? Recall that the energy amplication after each scattering event is A 4 2 = 16 3 kT mc2
2

Repeated scatterings for < 1

(8.56)

where the second equality holds only for a thermal electron distribution. Consider an initial photon distribution with mean energy much less than the rms electron energy i 2 1/2 mc2 with intensity I (i ). After k scatterings the energy of a photon of mean intial energy i is k i Ak . (8.57)

In the small limit the probability of the photon undergoing k scatterings is determined by the optical depth: p(k )sim k . Since the bandwidth of the Compton scattered spectrum is comparable to the frequency itself, the nal intensity at energy k has a spectrum: I (k ) I () k . (8.58)

This is expressed explicitly in terms of the initial and nal energies i and f respectively by writing k = Ak ln / ln A . One then has I (f ) = where = I (i ) 106 ln . ln A f i

(8.59) (8.60)

8.6

Kompaneets equation

Suppose that the Thomson optical depth is greater than unity and that the energy density in the photons is much greater than the energy density in the bulk of the electrons (ignoring any distribution of high-energy electrons). Then Compton scattering has two complementary eects. First, it causes the hard photons to be downscattered (due to the quantum recoil) to lower frequencies. Second, it causes soft photons to be upscattered (due to the Doppler eect) to higher frequencies. The energy transfer between the photons and electrons tends to determine the temperature of the electrons. Qualitatively one can understand these two energy transfer processes as follows. On the one hand, for an electron initially at rest, any recoil motion of the electron implies that the phtoon loses energy to the electron, and hence this must cause a downshift in the frequency of the photon and an energy transfer from photons to electrons. This quantum recoil eect dominates for thermal electrons. On the other hand, as a result of the Doppler eect, photons gain energy in head-on collisions with electrons and lose energy in overtaking collisions with electrons. This is due to there being a higher probability of a head-on collision (which causes an energy gain for the photon) than for an overtaking collision (which causes an energy loss for the photon). Hence, on average the photons gain energy from the electrons. Ignoring relativistic eects and for h me c2 , the average frequency downshift of hard photons due to the recoil is = h . m e c2 (8.61)

The frequency change due to the Doppler eect gives / v/c. The relative probability of head-on over overtaking scatterings implies that the net average upshift is / (v/c)2 . Averaging over a Maxwellian distribution gives the average frequency upshift for soft photons due to the Doppler eect = 4T . m e c2 (8.62)

Both these eects are included in a single equation, called the Kompaneets equation, which applies to an isotropic distribution of photons with occuption number N ( ) scattered by thermal electrons: T ne dN ( ) = dt me c 1 d 4 2 d T dN ( ) + N ( ) + [N ( )]2 . h d (8.63)

The rst term inside the curly brackets is the only one that involves the temperature of the electrons, and is describes the Doppler shift. The middle term describes the eect of the quantum recoil. The nal term describes the eect of induced scattering.

107

8.7

Questions

1. Determine the polarization pattern of Thomson-scattered radiation when the incident radiation is circularly polarized. Hint: write the incident waveeld as E = E0 ( x + iy) cos(t k0 x). (8.64)

2. Using (10.25), show that the total power scattered by the inverse Compton process is P = 4 3a 3a min T cWph K (3 a)1 max , 3 (8.65)

for a distribution of electrons N ( ) = K a , 0, min < < max . otherwise (8.66)

3. Show that the synchrotron half life time in the synchrotron-self-Compton model is t1/2 5 108 = B sin2 1G
2

seconds.

(8.67)

108

Chapter 9

Propagation Eects in Astrophysical Plasmas


Reference: Melrose & McPhedran, Chapter 10 Narayan, R. 1992, The Physics of Pulsar Scintillation, Phil. Trans. R. Soc. Lond. A, 341, 151165 Ishimaru, A., Wave propagation and scattering in random media, Wiley-IEEE Press
A medium with electromagnetic properties modies an electromagnetic eld imposed upon it. Thus radiation emitted by an astrophysical source is modied as it encounters plasma along the path toward an observer. Astrophysical plasmas lead to a variety of propagation eects that may be used to infer properties about interplanetary and intergalactic space. In this chapter we discuss a simple theory for the response of a plasma to electromagnetic radiation and then apply this to various astrophysical effects, including dispersion smearing, which is a frequency-dependent delay in the pulse arrival times of bursty-phenomena (such as the pulses from pulsars), and Faraday rotation, which rotates the plane of linearly polarized radiation. We conclude with a discussion of interstellar scintillation, which occurs when a wavefront propagates through an inhomogeneous medium, in this case the diuse ionised component of the interstellar medium. Interstellar scintillation gives rise to a variety of eects, particularly ux variability in compact radio objects such as pulsars, compact extragalactic sources and masers. In this chapter we make the dependence of various quantities on the permittivity and permeability of the media explicit.

109

9.1

The plasma response tensor

The response of a medium to a static uniform electromagnetic eld is described in terms of induced dipole moments. On a microscopic level a static uniform electric eld polarizes the atoms or molecules. The polarization, P is dened as the induced electric dipole moment per unit volume. For non-uniform static elds the response in terms of P remains approximately valid for any medium in which the multipole expansion converges suciently rapidly, as is the case for many media. In a similar way, the static response remains approximately valid when the assumption that the elds are static is relaxed to allow suciently slowly varying elds. It is conventional to introduce two additional elds when describing a static response. These are the electric induction, D and the magnetic eld strength, H, The electric induction is D := 0 E + P, H := B/0 M (9.1)

Traditionally for an isotropic medium one writes D = E and H = B/ where is the permittivity of the medium equal to 0 for free space and where is the permeability of the medium. The refractive index and the dielectric permittivity of the medium are related as follows 2 = c2 k 2 c2 k 2 = 2 . n (9.2)

For present purposes we are interested mainly in the eect of a medium on the electric induction. In general the response of the medium is not in the same direction as the electromagnetic disturbance, in which case the response of the medium is anisotropic and is replaced by the permittivity tensor, ij , or equivalently, t he dielectric tensor, Kij Di = ij Ej = 0 Kij Ej , The dielectric tensor is often written in the form Kij = = ij + i ij , 0 ij + ij (9.4) (9.5) (9.3)

where ij is known as the conductivity tensor and ij as the suspectibility tensor. Using the denition in (9.1) the polarization is Pi = 0 ij Ej . (9.6)

It is possible derive the response of a plasma to an electromagnetic wave using magnetoionic theory. This theory is simple in that the thermal motions of the electrons and ions in the plasma are neglected. The only charged particles assumed present in magnetoionic theory are electrons. It is usually assumed 110

that there is a uniform background charge density due to positive ions so that the plasma is charge-neutral, and that these ions play no role in the response. However, magnetoionic theory does take the eect of an ambient magnetic eld into account, and it reproduces th e results of the more general theory suciently well for our purposes. The electrons are treated as a continuous uid with number density ne and uid velocity v. The equation of uid motion is me dv(t) = e[E(t) + v(t) B] e me v, dt (9.7)

where the spatial dependence of E(t) is unimportant, and where B is the static magnetic eld. The nal term is a frictional drag assumed to be exerted on the electrons by ions, with e the electron-ion collision frequency. The uid velocity is replaced by v = dX(t)/dt, where X(t) is the position of the electron. The total time derivative in (9.7) is written d/dt = /t + v . When equation (9.7) is linearized the contribution of the convection term, v , is omitted and we Fourier transform the resulting equation to obtain me ( + ie )X( ) ieX( ) B = eE( ). (9.8)

The electromagnetic response of the plasma is described in terms of the Fourier transform of the polarization, P( ), which is the induced dipole moment per unit volume. It is identied as P( ) = ene X( ). (9.9)

If we introduce the unit vector b along the direction of the magnetic eld B, equation (9.8) can be written as
2 ( + ie )P( ) + i e P( ) b = 0 p E( ),

(9.10)

where e = eB , me p = (ne e2 /0 me )1/2 (9.11)

are the electron cyclotron frequency and the electron plasma frequency respectively. It is conventional to introduce the magnetoionic parameters
2 X = p / 2 ,

Y = e /,

Z = e /.

(9.12)

Then in a coordinate system in which the magnetic eld is along the z -axis one writes (9.10) as U iY 0 Px ( ) Ex ( ) iY U 0 Py ( ) = 0 X Ey ( ) . (9.13) 0 0 U Pz ( ) Ez ( ) 111

with

with U = 1 + iZ . On inverting (9.13) one obtains an equation of the form Pi ( ) = 0 ij Ej ( ) which allows us to identify the dielectric tensor as S iD 0 Kij ( ) = iD S 0 , (9.14) 0 0 P S =1 UX , Y2 D= XY , Y2 P =1 X . U (9.15)

U2

U2

Henceforth we neglect collisions, so one has Z = 0 and U = 1. In this case it may be shown that the natural modes of the medium are found by solving the matrix equation: 2 n 0 0 det Kij 0 n2 0 = 0, (9.16) 0 0 0 which has solutions1 P [(S n2 )2 D2 ] = 0. (9.17)

The solution P = 0 corresponds to longitudinal waves, and is of no interest. The transverse waves (i.e. waves with Ex and Ey both perpendicular to the direction of propagation) have solutions n2 = n2 = S D = 1 X XY. (9.18)

On substituting these two solutions into (9.16), one obtains the polarization vectors (9.19) e+ = eR = (1, i, 0)/ 2 e = eL = (1, i, 0)/ 2, where R and L denote right and left hand circular polarization. These two polarization vectors are the natural modes of the plasma. That is to say, right-hand circularly polarized radiation propagates through the plasma with a refractive index n+ , and left-hand circularly polarized radiation propagates through the plasma with refractive index n . Since any transverse eld can be decomposed as a linear combination of ER and EL as follows E = aR E R + aL E L , (9.20)

propagation of radiation of arbitrary polarization is treated by decomposing the radiation into a linear combination of right and left-hand circularly polarized modes.
1 Equivalently, we have could solved for the dielectric permittivities along each direction in the medium. The refractive and permittivity are related by n2 = /0 .

112

The generalisation to magnetic eld of arbitrary orientations is straightforward (see Qu. 1). If the wave frequency is much greater than e , the magnetic eld term in equation (9.10) has a small eect compared to the electric eld term. If we can ignore the motion of the electrons along the direction of wave propagation, one makes the replacement e e cos (9.21)

where is the angle between the magnetic eld and the direction of propagation of the radiation.

9.2

Dispersion smearing

For the moment, let us neglect the small dierence in the refractive indices of the two wave modes. Then the refractive index of both modes is 2 p n S 1 X/2 = 1 . (9.22) 2 2 This refractive index is less than unity, so the phase velocity /k is greater than c. However, the group velocity, vg =
2 p d =c 1 dk 2 2

(9.23)

determines the speed at which the signal travels. As can be seen from (9.23), this is less than c. The group velocity vg depends on both the frequency of the radiation and the plasma density. Low frequency radiation is delayed with respect to high frequency radiation as it travels through a plasma. The delay increases as the plasma density increases. For a signal travelling through a plasma of length L, the total time delay with respect to a signal propagation in vacuo is t = = L L vg c 2 L p , c 2 2

(9.24)

where we have assumed p in the second line. Thus the time delay is related to the medium density along the line of sight through the plasma. It is convenient to express the time delay as t = e2 2c 2
0 me

DM.

(9.25)

where DM is the dispersion measure and is dened as the integral of the electron density along the line of sight
L

DM

ne (z ) dz.
0

(9.26)

113

The SI units of the dispersion measure are m2 but in astronomy it is more commonly expressed in pc cm3 . The eect of this delay is observed in pulsars. One observes a pulsar signal over a range of frequencies. The same pulse is observed to arrive later at lower frequencies. Given the exact dierence in arrival times of the same pulse at dierent frequencies, it is possible to determine exactly how much plasma lies in between the pulsar and the telescope that observed the pulse. Given a mean electron density in the interstellar medium, it is then possible to roughly estimate the distance to a pulsar. The mean electron density in the interstellar medium is ne 0.03 cm3 .

9.3

Faraday rotation

We saw above that the refractive index of right and left-hand circularly polarized 2 radiation diers by an amount XY = p e cos / 3 . This gives rise to the eect of Faraday rotation, which rotates the plane of polarization of linearly polarized signal. To see this, consider a plane wave that is linearly polarized in the x direction at the source: eikzit . E = E0 x (9.27)

Now recall that the wave modes of the plasma are circularly polarized, namely e (1, i, 0). So we decompose the linearly polarized wave into two circularly polarized plane waves E= 1 ikzit ) + ( )] . e [( x + iy x iy 2 (9.28)

Now from (9.2) one writes the wavenumber for the two modes as k = k0 where we write k0 = c k =
2 p e cos , 2 2 c 2 p k 2c

(9.29)

(9.30)

for a magnetic eld that makes an angle with respect to the direction of propagation (cf. (9.21)). The wavenumber k0 is the value for propagation through free space. The rst correction in (9.29) is the correction due to the isotropic part of the plasma response tensor, and depends only on the electron density in the medium. The last correction term, k is the part that gives rise to Faraday rotation, and it depends on both the plasma density via its dependence on p and the magnetic eld via its dependence on e . After propagating a distance z through the plasma the electric eld is E= 1 ik0 zi )ei() + ( )ei(+) , ( x + iy x iy e 2 114 (9.31)

with
z

2 p dz 2c

k dz =
0 0

2 p e cos dz 2 2 c

(9.32)

The foregoing arguments can be used to treat the propagation of radiation of arbitrary polarization through a magnetized plasma. Faraday rotation does not alter the Stokes parameters I and V , but aects the linear polarization, by transforming initial Stokes paramters Q0 and U0 into the new Stokes parameters Q and U as follows Q U = cos 2 RM sin 2 RM sin 2 RM cos 2 RM
2 p e 2 8 c3

Q0 U0

(9.33)

where we have written the Faraday angle in terms of the rotation measure, RM: = 2 RM = 2 z . (9.34)

When Faraday rotation is observed, the rotation measure yields information on the magnetic eld and density uctuations along a given line of sight. In particular it is linearly proportional to the integral product of the electron density and the component of the magnetic eld along the line of sight along the ray path
z

RM

ne B dl.

(9.35)

Measurements of Faraday rotation have been used to deduce the large-scale magnetic eld structure of many astronomical objects, including our Galaxy, nearby spiral galaxies and the jets of radio galaxies. Faraday rotation measurements indicate that the mean magnetic eld in our Galaxy is B 3 G. The random component of the magnetic eld has an amplitude of comparable magnitude to the uniform component.

9.4

Interstellar Scintillation

The plasma in the interstellar medium is inhomogeneous. Fluctuations in the electron density introduce spatial variations in the phase across a wavefront. As the wavefront propagates, these phase distortions give rise to intensity uctuations across the surface of the wavefront. An object whose radiation is observed through an inhomogeneous medium exhibits temporal intensity variations provided several conditions are met. The objects ux density is observed to vary temporally (scintillate) if the medium moves relative to an observers line of sight to the source and provided the source is suciently compact. The twinkling of stars due to turbulent uctuations in the Earths atmosphere is a terrestrial manifestation example of this phenomenon of scinitillation. Interstellar scintillation occurs at radio frequencies due to random phase delays caused by inhomogeneities in the ionised component of the Galactic interstellar medium. 115

9.4.1

Propagation through an inhomogenous medium: The Parabolic Equation

We consider a wave E(r) propagating through a random medium with dielectric permittivity (r) = [1 + 1 (r)]. (9.36)

We have assumed that the medium is isotropic, so that the dielectric permittivity does not depend on the polarization of the wave. As we saw above, this is a good approximation to rst order. The term 1 (r) represents the spatial variations in the permittivity due to inhomogenieties in the medium. If we dene the average wavenumber 2 2 = k0 , c2 the wave equation describing the propagation of the electric eld E is k2 = 2 E + k 2 (1 + 1 (r)) E = 0. (9.37)

(9.38)

This is the scalar wave equation. We are able to treat the electric eld as a scalar in the present context because we have assumed that the dielectric permittivity does not depend on the polarization of the radiation (ie. we ignore eects associated with Faraday rotation). The rst step in deriving the parabolic equation is to assume a solution of the form E (r) = u(r)eikz , (9.39)

for a wave propagating along the z -axis. This approximation separates the quickly varying eikz term from the wave amplitude u(r) which is, under most cases of interest here, a slowly varying function of z . Substituting (9.39) into (9.38) we obtain u(r) + 2 u(r) + k 2 1 (r)u(r) = 0. z Since u(r) is a slowly varying function of z we note that 2ik k u 2u z z 2 (9.40)

(9.41)

provided that the scale size of the random medium is much larger than the wavelength. This is an excellent approximation in practice. The wave equation then reduces to the parabolic equation for u(r) 2ik u(r) 2 + 2 t u(r) + k 1 (r)u(r) = 0, z 2 2 + . 2 = t x2 y 2 (9.42) (9.43)

The parabolic approximation to the wave equation allows us to derive solutions for the average value of the waveeld and its higher moments. 116

Mutual coherence function The lowest order moment of interest in scintillation theory is the second moment, which is also known as the mutual coherence function (1 , 2 ; z ) = u(1 ; z )u (2 ; z ) . (9.44)

Physically, this quantity represents the correlation in the waveeld between two points, 1 and 2 located on some plane. We have written the dependence of z explicitly in the mutual coherence function to emphasise that the waves travel along the z axis and the plane on which 1 and 2 are measured is orthogonal to this axis. This is the quantity which is measured by any interferometer (e.g. the Westerbork radio telescope). It is directly related to the image brightness, B (x, y ) on the sky via a Fourier transform (1 , 2 ; z ) = z 2 d2 B (z ) exp [ik (1 2 ) ] . (9.45)

Thus if a source is compact B (x, y ) is sharply peaked. Its mutual coherence function is then a broad function of |1 2 |. For example, if B (x, y ) were a point source at the origin (ie. B (x, y ) = (x) (y )) the mutual coherence function would be at: (1 , 2 ; z ) = 1. The more extended the source brightness distribution on the plane of the sky, the more steeply peaked is the mutual coherence function about |1 2 | = 0. We now proceed to determine the average visibility of a scattered point source. We start with the parabolic equation (9.42) for u(1 ; z ) and multiply it by u (2 ; z ) 2ik u(1 ; z ) 2 u (2 ; z ) + 2 t1 u(1 ; z )u (2 ; z ) + k 1 (1 ; z )u(1 ; z )u (2 ; z ) = 0, z (9.46)

where 2 t1 is the Laplacian with respect to 1 . We can also take the conjugate of (9.42) with 1 replaced by 2 and multiply it by u(1 ; z ): 2ik u(2 ; z ) 2 u(1 ; z ) + 2 t2 u (2 ; z )u(1 ; z ) + k 1 (2 ; z )u (2 ; z )u(1 ; z ) = 0, z (9.47)

Subtracting (9.47) from (9.46) and taking the ensemble average (ie. a statistical average over all possible realizations of the random variables) yields 2 (1 , 2 ; z ) + (2 t1 + t2 )(1 , 2 ; z ) z +k 2 [1 (1 ; z ) 1 (2 ; z )]u(1 ; z )u (2 ; z ) = 0. 2ik

(9.48)

After some algebra and application of the Navikov-Furutsu formula (beyond the scope of this course), one may write the nal term in (9.48) in terms of 117

(1 , 2 ; z ) to obtain 2 (1 , 2 ; z ) + (2 t1 + t2 )(1 , 2 ; z ) z ik (0; z ) C (1 2 ; z )](1 , 2 ; z ) = 0. + [C 4 2ik

(9.49)

This is a dierential equation describing the propagation of the mutual coher ence function through a scattering medium. The quantity C () is the derivative (along the z -axis) of the phase autocorrelation function
z z

C () = (r + )(r) =
0

dz1 (r + 1 ; z )

dz1 (r; z )
0

. (9.50)

Each quantity in square brackets is the total phase due to scattering material along a path parallel to the z -axis. The correlation function is the correlation between the phase along a line through the medium parallel to the z -axis and ending at the point r + on the observers plane, and the phase along a line ending at the point r. We now illustrate the use of (9.49) with a specic solution. Suppose a wave from a point source located at innity is incident upon the scattering medium. This is represented by a plane wave whose mutual coherence function before entering the scattering medium is (1 , 2 ; 0) = 1. The solution of (9.49) is then 1 (1 2 ; z ) = exp D (1 2 ) , 2 where D (r) = 2[C (0) C (r)] = [(r + r) (r )]2 (9.53) (9.52) (9.51)

is the phase structure function, which is found to have the following form in the interstellar medium D (r) = r rdi
5/3

(9.54)

The quantity rdi is a characteristic scale length of the phase uctuations in the interstellar medium, over which the root mean square phase dierence is 1 radian. The stronger the scattering, the larger the phase variations in the medium over a given distance. Thus the quantity rdi decreases as the strength of the scattering increases. Depending on the actual line of sight and the observing frequency, rdi typically ranges from 104 m to 109 m. The mutual coherence function, or visibility, given by (9.52) has a simple physical interpretation: a point-like source acquires a nite angular diameter 118

due to scattering. The stronger the scattering, the greater the apparent angular diameter of the scattered source. Let us see how this comes about. First, consider the visibility in the limit of weak scatterng. Here, the phase uctuations are small, the quantity rdi tends to innity, and the phase structure function, D (r) is zero. Then the visibility of the source is (1 2 ; z ) = 1. From the relation between the visibility and the source brightness distribution (9.45), we see that this corresponds to the visibility for a point source. Now suppose that scattering is appreciable, and rdi has some nite value. Then the visibility begins to drop appreciably when the separation between receivers |1 2 | is comparable to rdi . Interpreted as a brightness distribution on the sky, the scattered source appears to subtend an angle = 1/krdi . In practice, scatter-broadening in the interstellar medium is only appreciable when (a) the initial (unscattered) source angular diamter is smaller than the scatter-broadening angle and (b) when the scattering is suciently strong that the apparent angular diameter of the source is suciently large. Scattering broadening is observed in pulsars and some extragalactic sources. It is also particularly important for Sgr A*, the nonthermal radio source at our Galaxys centre thought to b e associated with a 2.6 106 M black hole (see Fig 5.1). Intensity uctuations semi-quantitative Here we present a heuristic description of the intensity uctuations caused by scintillation. A more formal treatment is presented in the following section. Consider radiation from a point source at innity incident upon a phase screen which is located a distance D from an observer at z = zO , as shown in Fig. 5.2. Making the simplifying assumption that the phase screen is suciently thin that only phase modulation results at the exit plane of the screen, an incident wave with unit amplitude exits the screen with complex amplitude exp[i(x)], where (x) is the phase change introduced at the point x on the screen. The amplitude of the waveeld incident upon the phase screen, u(zO D; x), is then related to that on the observers plane, u(zO ; X), via the FresnelKircho integral u(zO ; X) = rF = i 2 2rF D/k, d2 x u(zO D; x) exp i(x X)2 + i(x) 2 2rF (9.55) (9.56)

where rF is the Fresnel scale and k is the wavenumber. It is assumed that the phase uctuations are frozen onto the scattering screen, so that any temporal uctuations in the waveeld at X are due to relative movement between the screen and the source-observer line of sight. The thin-screen approximation is used extensively in astrophysics, despite the fact that the radiation is likely to propagate through many turbulent regions distributed along the line of sight. However, the results discussed in the context of this model dier only slightly with the path-integral approaches used

119

plot of measured (FWHM) source size vs. observing A* (1997 February 714). The solid line represents a 1.43 t e axis sizes (open circles), while the dashed line a 0.76 l2.0 t to the minor axis sizes ( lled circles).

Figure 9.1: Radiation from the black hole candidate at the Galactic Centre, Sgr A*, is heavily scatter broadened by interstellar scintillation. The log-log plot above shows the measured (FWHM) source size of Sgr A* vs. observing wavlength. The solid line r epresents a t to the major axis sizes (open circles), while the dashed line shows a t to the minor axis sizes (lled circles). (From Lo et al. 1998.)

120

(x)

S
(at infinity)

observer's plane z=zO

phase screen z=zO -D


Figure 9.2: Geometry of the scattering for the thin-screen approximation. to model scattering in extended inhomogeneous media. In particular, propagation through an extended medium alters the degree of refractive scintillation expected (see 9.4.3 below). The following discussion is restricted to discussion of the thin-screen approximation only. Wavefront coherence Equation (9.55) implies that two terms contribute to the coherence of the wavefront propagating through an inhomogeneous medium: distortions due to the inhomogeneities, characterized by (x), and geometric path length variations, 2 characterized by the term (x X)2 /2rF . Neglecting phase variations due to inhomogeneities on the screen, one sees that geometric path length variations truncate the integral as |x X| becomes comparable to rF , causing the exponential term to vary quickly. The observed wave amplitude is therefore dominated by the radiation within a circle of radius rF on the phase screen, known as the Fresnel zone. The physical interpretation is that radiation emanating from the region outside the Fresnel zone wraps in geometric phase quickly, and does not contribute coherently to the wave amplitude at X.

121

The eect of density uctuations in the ISM is embodied in the phase delay term, (x). The statistics of the density uctuations are assumed to be widesense stationary meaning that (x) is independent of x and that (x1 )(x2 ) is only dependent on the quantity x1 x2 . It is further assumed, by the central limit theorem, that is a random gaussian variable. Since the second moment of a gaussian random variable is sucient to describe all its higher moments, the phase structure function, D (r) = [(r + r ) (r )]2 , (9.57)

completely describes the statistics of the phase uctuations on the screen. The angular brackets in (9.57) denote an ensemble average of the random variable over all possible realizations of the scattering screen. For many lines of sight through the ISM the phase structure function reects an underlying power law spectrum of phase inhomogeneities between some inner and outer scales l0 and L0 respectively, and one has D (r) = r rdi
2

(9.58)

where is close to 11/3, the value expected for Kolmogorov turbulence. The strength of the phase uctuations on the screen is characterized by the diractive scale, rdi , over which the root-mean-square phase dierence on the phase screen is one radian. The relative importance of the geometric and screen phase terms is determined by the relative magnitude of the Fresnel and diractive scales. Two limiting cases are considered below: weak scattering, corresponding to rdi rF , and strong scattering, in which screen phase dominates geometric phase, rdi rF . The scattering strength depends on the observing frequency, since both rF and rdi 2/( 2) are frequency dependent. The scattering strength decreases with frequency. These two regimes are discussed further in 5.4.3.

9.4.2

Intensity uctuations quantitative

We dene the fourth moment of the waveeld as the ensemble average correlation in the waveelds measured by four receivers at the points 1 , 1 , 2 and 2 :
4 = u(1 )u ( 1 )u(2 )u (2 ) .

(9.59)

It is convenient to transform these co-ordinates into the new co-ordinates 1 ( + 2 + 1 + 2 ) = 1 + 2 1 2 4 1 1 1 r1 = (1 2 + 1 2 ) r2 = (1 2 1 + 2 ) 2 2 R= (9.60) (9.61)

Using arguments similar to those in deriving (9.49) it is possible to construct an expression for the propagation of the fourth moment of the waveeld, which 122

r1 u*(x4) u(x1)

r2

u(x3)

u*(x2)

Figure 9.3: Conguration of the four receivers in (9.64) for = 0. can be written in terms of our new coordinates as i (R + r1 r2 ) + Q 4 = 0 z k where Q =
D (r1 + /2) + D (r1 /2) + D (r2 + /2)

(9.62)

+ D (r2 /2) D (r1 + r2 ) D (r1 r2 ).

(9.63)

Now consider a plane wave propagating in a random medium. This wave has no dependence on the centre co-ordinate R, so we have R = 0. Since is also eliminated from the equation, is only a parameter and we are free to set it to zero. In this case the four co-ordinates 1 , 1 , 2 and 2 lie on the vertices of a parallelogram. We write (9.62) as i r1 r2 4 = 0. z k (9.64)

We can solve (9.64) for a thin screen. We outline the solution here. Assume we have a plane wave incident upon the scattering screen. If write 4 = exp[ ] equation (9.64) implies i [r2 r1 + r2 r1 ] Q = 0 z k Integrating w.r.t z we have = Qz + i k
z 0

(9.65)

[r2 r1 + r2 r1 ] dz.

(9.66)

For a screen of thickness z we have, to rst order, = Qz. 123 (9.67)

Thus the fourth moment of the waveeld upon exiting the scattering screen is 4 = exp( ) (9.68) = exp[D (r1 ) D (r2 ) + D (r1 + r2 )/2 + D (r1 r2 )/2]. (9.69)

Now we consider propagation of this fourth moment in free space from z = 0, where the thin screen is located, to an observer located at z = D. For propagation in free space one sets Q = 0 in equation (9.64). Solving this equation with initial condition set by (9.69) one obtains 4 (r1 , r2 ; z ) = k 2z
2 dr 1 dr2 4 (r1 , r2 0) exp ik (r1 r 1 ) (r2 r2 ) z

1 exp {2D (r 1 ) + 2D (r2 ) + D (r1 + r2 ) + D (r1 r2 )} . 2 (9.70) Henceforth we assume that the incident wavefront is from a point source located at innity of unity intensity, so one has 4 (r1 , r2 ; 0) = 1. We can derive the characteristics of the intensity uctuations by setting r2 = 0 in (9.70). One then has 4 (r1 , 0) = I (r + r1 )I (r) .

9.4.3

Scintillation regimes

Weak scattering Weak scattering refers to the scintillation regime in which the phase uctuations due to inhomogeneities in the ISM are small (i.e. less than one radian) on length scales smaller than the Fresnel scale: rdi rF . Radiation from a pointlike source appears scattered over a region of radius rF . Slight curvature in the wavefront is induced by curvature in the phase delay caused by the inhomogeneities. The radiation is focussed and defocussed due to this curvature and, as a result, the source is observed to undergo ux variability about the mean ux density of the source. In the limit rdi rF , the intensity modulation index, m [I I ]2 1/2 / I , is approximately (rF /rdi )/21 < 1. This result may be understood in terms of geometric optics. The phase inhomogeneities over the scattering disk act like a weak lens whose focal length is determined by root-mean-square phase change across the disk, D (rF )1/2 = (rF /rdi )/21 . The amplitude of the fractional variability is estimated by D/f , where f 2 rF /D (rF )1/2 is the focal length2 of the lens. The ux density of a source varies as dierent patches of the phase screen move across the line of sight, and the variability timescale is approximately the Fresnel scale divided by a velocity characteristic of the scintillation; this is termed the Fresnel timescale: tF = rF /v .
2 The focal length of a parabolic lens which retards the wavefront by a length /k over a length scale r is r 2 k/.

124

In the weak uctuation limit one expands the last of the exponential terms in (9.70), to rst order in D(r) to obtain, after some manipulation, 4 (r1 , 0; z ) = 1 + 4 (2 )2 d2 q (q) exp[ir1 q] sin2 q2 z 2k , (9.71)

where (q) is the power spectrum of the phase uctuations and is related to the phase structure function as follows D (r) = 2 d2 q {1 exp[iq r]} (q). (2 )2 (9.72)

Equation (9.71) is often called the Born approximation to the intensity uctuations. For phase uctuations following a power spectrum (q) q it is possible to obtain closed-form expressions for the intensity autocorrelation function. However, as the scattering strength increases (ie. rdi decreases) the accuracy of the Born approximation diminishes. The Born approximation is invalid for scattering strengths rdi < rF = z/k. Another approximation is necessary to derive the character of the intensity uctuations for strong scattering, with rdi rF . Strong scattering The scintillation enters the regime of strong scattering when the screen phase uctuations dominate the geometric phase on the Fresnel scale (i.e. rdi rF ). Since the screen phase uctuations dominate the geometric phase, regions well outside the Fresnel zone contribute coherently to the integral in equation (9.70). This occurs at points at which the screen phase osets the contribution of the geometric phase such that the total phase term in (9.55) is approximately con2 stant (i.e. d(x)/dx = d[(x + X)2 /2rF ]/dx). Each of these coherent patches is of radius rdi , corresponding to the scale length over which the phase is approximately constant. Flux variability in the strong scattering regime occurs due to two distinct physical processes provided < 4, which is the case of interest here. The rst, termed diractive scintillation, is caused by the twinkling on and o of the coherent patches. In the stationary phase approximation (see Born & Wolf 1965), the waveeld at the observers plane is modelled as a sum of contributions from the coherent patches (points of stationary phase) on the phase screen. The contributions from each stationary phase point are of similar amplitude but random phase. Fluctuations in the observed waveeld are due to interference between the numerous coherent patches; this results in an observed waveeld whose amplitude and phase obey the statistics of a twodimensional walk on the complex plane. This results in intensity uctuations whose probability distribution function follows a negative exponential form, p(I ) = (1/I0 ) exp(I/I0 ), where I0 is the mean source intensity. A source undergoing diractive scintillation exhibits uctuations with modulation index mdi = 1 about the mean intensity of the source. Diractive ux variations 125

occur on the timescale comparable to that required for a coherent patch to drift across the line of sight, tdi = rdi /v . The second type of ux variability, refractive scintillation, arises due to large scale phase curvature across the phase screen, which regulates the number of coherent patches participating in the scintillation. Consider the area over which patches are visible on the screen. Each coherent patch acts like a narrow diracting aperture of size rdi which scatters radiation into a cone of angle 1/k rdi (e.g. Narayan 1992). An observer looking toward the phase screen sees radiation from patches out to this angle, and the apparent radius of this envelope of 2 coherent patches is termed the refractive scale, rref rF /rdi . A point-source located behind the scattering screen appears scatter-broadened out to an angle rref /D = 1/krdi . Refractive intensity uctuations are associated with phase curvature across the phase screen on a scale rref . This phase curvature induces weak refractive focussing and defocussing of the radiation, and regulates the number of diractive patches visible to an observer from the outer edges of the scattering disk. The ux increases or decreases as a result. Refractive ux variations occur on a timescale tref = rref /v , corresponding to the timescale over which the phase curvature changes appreciably across the scattering disk, of size rref . The nature of the refractive scintillation renders it amenable to modelling in terms of geometric optics. The modulation index due to refractive scintillation, mref = (rdi /rF )4 , < 4, may be derived in terms of geometric optics. As in the case of weak scattering, the modulation index is given by D/f , where the 2 /D (rref )1/2 . focal length of the lens due to the phase curvature is f = rref One can derive the characteristics of both diractive and refractive scintillation directly from (9.70). We will demonstrate this by deriving an expression for the spatial decorrelation of the intensity uctuations due to diractive scintillation. One can expand (9.70) as follows 4 (r1 , 0; z )
(r1 r k 1 ) (r2 ) dr 1 dr2 exp ik 2z z [exp [D (r1 )] (1 {D (r2 ) D (r 1 + r2 ) D (r1 r2 )}) 2

+ exp [D (r 2 )] (1 {D (r1 ) D (r1 + r2 ) D (r1 r2 )})] . (9.73)

If we neglect the terms in curly brackets these are associated with refractive scintillation one can perform the integral over r 2 by recognising that the term exp[D (r1 )] 1 is independent of r 2 and the term exp[D (r2 )] 1 is independent of r 1 . The contribution from diractive scintillation is then 4 (r1 , 0; z )di =
dr 1 (r1 r1 ) exp[D (r1 )] + dr 2 (r2 ) exp[D (r2 )]

= 1 + exp[D (r1 )].

(9.74)

Thus the contribution from diractive scintillation has I 2 = 2, which implies

126

(a)

(b)

rdiff

yyyyyyy @@@@@@@ @@@@@@@ yyyyyyy @@@@@@@ yyyyyyy @@@@@@@ yyyyyyy

O O

phase screen phase front

Figure 9.4: (a) Each diractive patch acts like a slit of size rdi scattering radiation into a cone of opening angle 1/k rdi . An observer therefore sees diractive patches across a distance D/krdi = rref on the scattering screen. (b) Any additional phase curvature on the screen due to large-scale irregularities causes focussing/defocussing of the radiation, and additional/fewer diractive patches become visible to an observer.

127

a modulation index of unity: ]2 = I 2 I = 1. [I I (9.75)

As can be seen from (9.74), the intensity uctuations due to diractive scintillation decorrelate as the function D (r1 ) becomes comparable to unity. This occurs on a spatial scale, r1 = rdi . Hence diractive scinitillations decorrelate on a spatial scale rdi , the length scale over which the rms phase change on the scattering screen is unity. An observer whose line of sight to a source moves at a speed v relative to the scattering medium percieves temporal variability due to diractive scintillation on a timescale rdi /v . The characteristics of refractive scinitillation are derived from the terms in curly brackets in (9.73).

9.4.4

Scintillation of non point-like (extended) sources

It is well known that stars twinkle at optical wavelengths due to atmospheric turbulence but that planets do not because of their larger apparent angular size. Source size considerations are equally important for sources undergoing interstellar scintillation. A source of nite size exhibits a reduction in the modulation index and increase in the timescale of variability relative to a point source. These eects are particularly applicable to a source whose angular extent, S , considerably exceeds the angular extent of the scintillation pattern F = rF /D in the weak scattering regime, or di = rdi /D or ref = rref /D in the strong scattering regime. Figure 9.4.4 illustrates how an extended source quenches the eect of scintillation. Consider the waves emitted from two point-sources, each of intensity I0 /2, on either side of an extended source of angular extent S . The waves propagate towards an observer with wavevectors whose angle diers by S . The two waves therefore intersect the screen at slightly dierent angles, and the complex wave amplitudes of the two wavefronts incident upon the phase screen, u1 (zO D; r) and u2 (zO D; r), say, are dierent. Equation (9.55) then implies that the observed waveelds of the two sources, u1 (zO ; X) and u2 (zO ; X), are also dierent. An observer therefore sees the sum of two partially independent uctuating signals. The partial decorrelation between the two signals implies that the variance of the resulting sum is less than the variance of a point source of intensity I0 . The timescale of uctuations also increases as some of the ne-scale scintillation structure is averaged out. The loss of correlation between the two wavefronts depends upon the relative separation, S D, of the apparent image centres of the two waves incident on the phase screen. Signicant decorrelation occurs when this separation becomes comparable to the length scale over which intensity uctuations result: rF in the weak scattering regime, and rdi and rref for strong scattering. Finite source size eects are therefore applicable for source sizes S > F in the weak scattering regime, and S > di for strong diractive scattering and S D > ref for strong refractive scattering.

128

S2 s S1
Ds

I1

I2

extended source

thin phase screen

Figure 9.5: The observed intensity scintillation patterns, I1 (x) and I2 (x), from opposite points S1 and S2 respectively, dier. For a source of 2-dimensional angular size S , the two proles are related by I1 (x) = I2 (x + DS ). The spatial decorrelation of the scintillation pattern for a point source is then used to compute decorrelation between I1 (x) and I2 (x) and hence the reduction in modulation index due to nite source size.

129

The foregoing discussion is restricted to two point-like sub-components of an source. However, these arguments are applicable to sources of arbitrary brightness distribution. In particular, the generalization to an arbitrary extended source is made by dividing the source into a number of point sources. The decrease in modulation index is derived by calculating the decorrelation between the waveelds of all pairs of sub-sources. At radio wavelengths the only sources small enough to exhibit ISS are pulsars, compact extragalactic sources and possibly some masers. Source size effects are particularly important when considering the variability of extragalactic sources, and may be used to determine a sources angular diameter. Only pulsars are suciently small to exhibit diractive scintillation. Indeed, even for these objects, the scattering may be suciently strong along some Galactic lines of sight at low frequency that source size eects become important. On the other hand, the requirement that a source exhibit refractive scintillation, S > ref , is more easily satised as the scattering becomes stronger (i.e. as rdi /rF decreases). Extended sources may therefore show greater variability at low frequencies where the scattering is stronger even though the refractive scintillation modulation index of a point source is smaller. Brightness temperatures of Intra-day variable radio sources The short (124 hr) timescales observed in some extragalactic radio sources at frequencies 5 10 GHz are attributed to scintillation in the regime of weak scattering. Extragalactic objects are chiey observed along lines of sight o the Galactic plane. There is considerably less scattering material relative to lines of sight through the Galactic plane, and strong scattering only occurs at frequencies below 4 7 GHz. The asymptotic theory of scintillation in the weak regime (i.e. scintillation theory in the limit rdi rF ) is used to derive the angular sizes of these sources, despite the fact that the theory is not strictly valid for the regime rdi rF applicable to these sources. Nonetheless, the hourly to daily variability timescales observed are used to derive angular diameters in the range 5 50 as. The exact r ole of scintillation in the variability of extragalactic sources is controversial. The small angular diameters implied by the rapid variability leads, in some cases, to sources whose brightness temperatures are dicult to reconcile with the present understanding of the radiative processes in these sources.

9.5

Questions

1. The question is concerned with Faraday rotation when the magnetic eld makes an angle to the direction of propagation. For propagation when the magnetic eld makes an angle to the z -axis in the xz plane equation

130

(9.13) becomes U iY cos iY cos U 0 iY sin

with

One then nds following dielectric tensor corresponding to (9.14) S B sin2 iD cos B sin cos , Kij = iD cos (9.77) S iD sin 2 B sin cos iD sin S B cos B= Y2 . UY 2 (9.78)

0 Px ( ) Ex ( ) iY sin Py ( ) = 0 X Ey ( ) (9.76) . U Pz ( ) Ez ( )

U3

If we neglect collisions and neglect terms second order in Y 2 we have S iD cos 0 S iD sin , Kij = iD cos (9.79) 0 iD sin S with S = 1 X, D = XY. (9.80)

Show that natural wave modes are still given by equation (9.19) with the refractive indices
n2 = S D cos .

(9.81)

2. Derive (9.70) from (9.64) under the assumptions outlined in the text. Hint: in dealing with the propagation of 4 from z = 0 to z = L use i r1 r2 z k 4 = 0 (9.82)

and take the Fourier transform of 4 (r1 , r2 ) with respect to r1 and r2 to write M (1 , 2 ) = to obtain i 1 2 M (1 , 2 ) = 0. z k Solve this for M and Fourier transform back to obtain (9.70). (9.84) 1 (2 )4 dr1 dr2 4 (r1 , r2 ) exp[i1 r1 i2 r2 ], (9.83)

131

Chapter 10

Exercises
Radiation Astrophysics 2010
Set of Exercises for the Course
By the end of the course, you should be able to do all of these homework exercises. The exercises come from a variety of sources, including Rybicki & Lightman, and the minds of Jean-Pierre Macquart and Penny Sackett. They vary in diculty and are in random order. Quickly read all of the problems now. Reread them again after each and every lecture, marking those which you feel able to solve. Solve these as quickly as you can after each lecture; you are encouraged to work together and share knowledge. Each week you will be asked to solve at least one of the problems on this list. You will not be told in advance which exercise will be requested, but it will be one for which you have already heard the relevant lecture material. The instructor will sketch the correct solution immediately thereafter. There is, therefore, no possibility to repeat or postpone this weekly process.

132

1.

Show that an interstellar grain in thermal equilibrium with gas at T 100K rotates rapidly. If its radius is a 105 cm, and its density is similar to that of water, show that the angular velocity is 105 to 106 rad s1 . A particle of mass m and charge e moves at constant, nonrelativistic speed v1 in a circle of radius a. a) What is the power emitted per unit solid angle in a direction at angle to the axis of the circle? b) Describe qualitatively and quantitatively the polarization of the radiation as a function of the angle . c) What is the spectrum of the emitted radiation? d) Suppose a particle is moving nonrelativistically in a constant magnetic eld B. Show that the frequency of the circular motion is B = eB/mc, and that the total emitted power is 2 2 P = ro c(v /c)2 B 2 , 3 and is emitted solely at the frequency B . (The nonrelavistic form of synchrotron radiation is called cyclotron or gyro radiation). The life time of an electron in the rst and second excited states of hydrogen is about t = 108 s. a) What is the natural broadening of the hydrogen H at = 6563 A, which is a transition between these two states? b) This line is created copiously in the photosphere of the Sun, where the line is Doppler broadened. Estimate the width of the H line in the Suns photosphere, and compare it to the natural width of the line. c) The solar H line is also pressure broadened. Estimate the size of the pressure broadening assuming a number density of hydrogen atoms of about 1.5 1017 cm3 . Compare to the natural width of the line.

2.

3.

133

4. a) An ultrarelativistic electron emits synchrotron radiation. Show that its energy decreases with time according to = 0 (1 + A0 t)1 where
2 2e4 B . 3 m 3 c5 Here 0 is the initial value of and B = B sin .

A=

b) Show that the time for the electron to lose half of its energy is t1/2 = (A0 )1 = 5.1 108 2 0 B .

5. Download the spectral data for your personal star from the Astrofysica webpage. The rst column is wavelength (given in nm) and the second column is ux per unit wavelength (uncalibrated units). a) Plot F vs of the data twice, rst on a linear-linear plot, then on a log-log. b) Repeat a) but plotting F vs . c) Estimate the temperature of the atmospheric layer of the star that emitted this radiation by tting a blackbody spectrum to your stellar spectrum. You might consider writing a simple supermongo program, for example, to plot a blackbody spectrum on top of your data, adjusting the free parameters until the two look similar. Remember that the stellar ux is uncalibrated. What would you give as the uncertainty (error bar) on your temperature estimate. Why? d) Can you use this determine the stellar type uniquely? Why or why not? 6. Show that the Larmor radius of a proton, moving at 10 km sec1 through a eld of 106 gauss, is small compared to interstellar and even interplanetary distances. 7. Photoionization is a process in which a photon is absorbed by an atom (or molecule) and an electron is ejected. An energy at least equal to the ionization potential is required. Let this energy be ho and let be the cross section for photoionization. Show that the number of photoionizations per unit volume and per unit time is

4na
o

J d = cna h

u d, h

(10.1)

where na = number density of atoms. 8. Single pulses from a pulsar are recorded at a telescope with bandwidth of 100 MHz and centre observing frequency 1200 MHz. The pulse arrival time diers by 48.7 ms from one end of the band to the other. What is the dispersion measure of this pulsar? Assuming the average electron density in the interstellar medium is ne = 0.03 cm2 , estimate the distance to the pulsar. 134

9. a) Show that the condition that an optically thin cloud of material can be ejected by radiation pressure from a nearby luminous object is that the mass to luminosity ratio (M/L) for the object be less that /(4Gc), where G = gravitational constant, c = speed of light, = mass absorption coecient of the cloud material (assumed independent of frequency). b) Calculate the terminal velocity v attained by such a cloud under radiation and gravitational forces alone, if it starts from rest a distance R from the object. Show that v2 = 2GM R L 1 4GM c (10.2)

c) A minimum value for may be estimated for pure hydrogen as that due to Thomson scattering o free electrons, when the hydrogen is completely ionized. The Thomson cross section is T = 6.651025 cm2 . The mass scattering coecient is therefore > T /mH , where mH = mass of hydrogen atom. Show that the maximum luminosity that a central mass M can have and still not spontaneously eject hydrogen by radiation pressure is LEDD = 4GM cmH /T = 1.251038 erg s1 (M/M ), (10.3) where M = mass of sun = 21033 g. 10. Show that if stimulated emission is neglected, leaving only two Einstein coecients, an appropriate relation between the coecients will be consistent with thermal equilibrium between the atom and a radiation eld of a Wien spectrum, but not of a Planck spectrum. 11. A certain gas emits thermally at the rate P ( ) (power per unit volume and frequency range). A spherical cloud of this gas has radius R, temperature T and is a distance d from earth (d >> R). a) Assume that the cloud is optically thin. What is the brightness of the cloud as measured on earth? Give your answer as a function of distance b away from the cloud center, assuming the cloud may be viewed along parallel rays as shown in the gure. b) What is the eective temperature of the cloud? c) What is the ux density F measured at earth coming from the entire cloud? d) How do the measured brightness temperatures compare with the clouds temperature? e) Answer parts (a)-(d) for an optically thick cloud. 12. Many astronomical objects are nearly spherical objects with a temperature gradient; layers of material at dierent temperatures surround a central core like the shells of an onion. Consider a spherical, opaque object emitting radiation as a blackbody at temperature TC . Surrounding this central object is a spherical shell of material which is emitting thermal radiation at a smaller temperature TS (TS <TC ). Suppose this shell absorbs in a narrow spectral line located at frequency 0 . (In other words, 0 >> 1 for other frequencies 1 close to 0 .) An astronomer gathers light from the object at frequencies 0 and 1 from two dierent rays: Ray A coming directly from the center of the object, and a parallel Ray B passing just through the outer shell. a) Will the observed brightness be larger at 0 or 1 for Ray A? For Ray B? b) Would your answers change if (TC <TS ) ? 135

13.

In certain cases the process of absorption of radiation can be treated by means of the macroscopic Maxwell equations. For example, suppose we have a conducting medium, so that the current density j is related to the electric eld E by Ohms Law, j= E, where is the conductivity (cgs unit = sec1 ). Investigate the propagation of electromagnetic waves in such a medium and show that: a) The wave vector k is complex, with k2 = 2 m2 /c2 , where m is the complex index of refraction, dened by 4i m2 = 1 + b) The waves are attenuated as they propagate, corresponding to an absorption coecient = 2 Im(m) c

(Note: In some literature, minus signs appear in these formulas. This is because the wave is often taken to be exp(kr+it) rather than the exp(krit) chosen here.) 14. Spatial uctuations in the plasma electron density in the interstellar medium give rise to twinkling of radiation from pulsars and other compact objects at radio wavelengths. A general expression for the correlation in the intensity uctuations at points r and r1 + r on the observers plane is I (r1 + r)I (r) = k 2z
2 2 d2 r 1 d r2 exp ik

(r 1 r1 ) r2 exp[G (r 1 , r2 )], (10.4) z

G (r 1 , r2 ) = D (r1 ) + D (r2 ) D (r1 + r2 )/2 D (r1 r2 )/2, (10.5)

where z is the eective distance to the scattering medium, k is the wavenumber, and the phase structure function, D is dened in terms of the power spectrum of phase uctuations, (q) as follows: D (r) = 2 d2 q [1 eiqr ](q). (2 )2 (10.6)

a) In the limit of weak scattering the phase uctuations are small and one expands (10.4) by writing exp[G ] = 1 G . Show that in this limit the correlation in the intensity uctuations may be expressed in the form I (r1 + r)I (r) = 4 (2 )2 d2 q exp[ir1 q] sin2 (q 2 z/2k ) (q). (10.7)

b) Argue that in the limit of weak scattering there is a characteristic spatial scale to the intensity pattern, and that intensity uctuations occur primarily on a spatial scale r1 z/k. [ Hint: the exact form of (q) is unimportant; it only suces to know that it is a sharply decreasing function of |q|.]

136

15.

A pulsar is conventionally believed to be a rotating neutron star. such a star is likely to have a strong magnetic eld, Bo , since it traps lines of force during its collapse. If the magnetic axis of the neutron star does not line up with the rotation axis, there will be magnetic dipole radiation from the time-changing magnetic dipole, m(t). Assume that the mass and radius of the neutron star are M and R, respectively; that the angle between the magnetic and rotation axes is ; and that the rotational angular velocity is . a) Find an expression for the radiated power P in terms of , R, Bo and . b) Assuming that the rotational energy of the pulsar is the ultimate source of the radiated power, nd an expression for the slow-down time scale of the pulsar. c) For M = 1M = 21033 g, R = 106 cm, Bo = 1012 gauss, = 90 nd P and for = 104 s1 , 103 s1 , 102 s1 . (The highest rate, = 104 s1 , is believed to result from rejuvenated pulsars in binary systems, where mass transfer adds angular momentum to an old pulsar, thereby spinning it up to very high rotation rates.)

137

16.

The Stokes parameters of a quasi-monochromatic plane wave propagating along the z -axis are dened in terms of the electric elds along the x and y -axes as follows I =
Ex Ex + Ey Ey Ex Ex Ey Ey Ex Ey + Ex Ey i( Ex Ey Ex Ey ).

Q = U = V =

(10.8)

a) Show that a plane wave with electric eld oscillating as ]/ 2 ER = eikzit [ x + iy

(10.9)

corresponds to right-hand circularly polarized radiation. Derive an expression similar to (10.9) for left-hand circularly polarized radiation EL . b) The electric eld for radiation that is linearly polarized along the x-axis propagates according to eikzit . E1 = x (10.10)

Decompose this electric eld into right and left-hand circularly polarized components. ie. write E1 in the form E 1 = aR E R + aL E L , (10.11)

and nd the coecients aR and aL . c) Faraday rotation is an eect in which a magnetized plasma introduces a phase dierence, , between right- and left-hand circularly polarized radiation as follows ER EL ER ei EL e .
i

(10.12) (10.13)

Show that radiation initially linearly polarized in the x direction, as in (10.10), is converted into radiation with an electric eld sin ) E = eikzit ( x cos + y (10.14)

after Faraday rotation. d) Using the denitions of the Stokes parameters, show that Faraday rotation in a magnetized plasma transforms initial Stokes parameters (I0 , Q0 , U0 , V0 ) into new Stokes parameters (I, Q, U, V ) as follows I0 1 0 0 0 I Q 0 cos sin 0 Q0 (10.15) U = 0 sin cos 0 U0 . V0 0 0 0 1 V

138

17. An optically thin cloud surrounding a luminous object is estimated to be 1 pc in radius and to consist of ionized plasma. Assume that electron scattering is the only important extinction mechanism and that the luminous object emits unpolarized radiation. a) If the cloud is unresolved (angular size smaller than angular resolution of detector), what is the net polarization observed? b) If the cloud is resolved, what is the polarization direction of the observed radiation as a function of position on the sky? Assume only single scattering occurs. c) If the central object is clearly seen, what is an upper bound for the electron density of the cloud, assuming that the cloud is homogeneous? 18. A plane-polarized wave is incident on a sphere of radius a, composed of a solid material. We assume that the wavelength is large compared with a. In that case, it is known that the electric eld at any instant of time is constant throughout the sphere and has the value E = E/(1 + 4/3), where E is the external (applied) eld and is the polarizability of the material. The dipole moment per unit volume is simply proportional to the internal electric eld P = E . Show that the total cross section for scattering the radiation is = a2 Qscatt where 8(ka)4 Qscatt = . 3(1 + 3/4)2 19. Consider a medium containing a large number of radiating particles. (For deniteness, you may wish to imagine electrons emitting Bremsstrahlung.) Each particle emits a pulse of radiation with an electric eld Eo (t) as a function of time. An observer will detect a series of such pulses, all with the same shape but with arrival times t1 , t2 , t3 , ..., tN . The measured electric eld will be
N

E (t) =
i=1

Eo (t ti )

a) Show that the Fourier transform of E (t) is


N

( ) = E o ( ) E
i=1

eit ,

o ( ) is the Fourier transform of Eo (t). where E b) Argue that


N

i=1

eit |2 = N

when averaged of the random arrival times. c) Thus show that the measured spectrum is simply N times the spectrum of an individual pulse. (Note that this result still holds if the pulses overlap.) d) By contrast, show that if all the particles are in a region much smaller than a wavelength and then emit their pulses simultaneously, then the measured spectrum will be N 2 times the spectrum of an individual pulse.

139

20.

In astrophysics it is frequently argued that a source of radiation which undergoes a uctuation of duration t must have a physical diameter of order D ct. This argument is based on the fact that even if all portions of the source undergo a disturbance at the same instant and for an innitesimal period of time, the resulting signal at the observer will be smeared out over a time interval tmin D/c because of the nite light travel time across the source. Suppose, however, that the source is an optically thick spherical shell of radius R(t) that is expanding with relativistic velocity 1, >> 1 and energized by a stationary point at its center. By consideration of relativistic beaming eects show that if the observer sees a uctuation from the shell of duration t at time t, the source may actually be of radius R < 2 2 ct , rather than the much smaller limit given by the nonrelativistic considerations. In the rest frame of the shell surface, each surface element may be treated as an isotropic emitter. (The argument has been used to show that the active regions in quasars may be much larger than ct 1 light month across, and thus avoids much energy being crammed into so small a volume.

21.

Show than an observer moving with respect to a blackbody eld of temperature T will see blackbody radiation with a temperature that depends on angle according to T = (1 v 2 /c2 )1/2 T 1 (v/c) cos

a) The 2.7K universal blackbody radiation at = 3 cm is isotropic to about one part in 103 . What is the maximum velocity that the earth can have with respect to the frame in which the radiation is isotropic? [Isotropy is measured by the ratio (Imax Imin )/(Imax + Imin ).] 22. A particle is accelerated by a force having components F|| and F with respect to the particles velocity. Show that the radiated power is
2 2 P = (2e2 /3m2 c3 )(F|| + 2 F )

Thus the perpendicular component has more eect in producing radiation than the parallel component by a factor of 2 .

140

23. Consider a sphere of ionized hydrogen plasma that is undergoing spherical gravitational collapse. The sphere is held at constant isothermal temperature T0 , uniform density and constant mass M0 during the collapse, and has decreasing radius R(t). The sphere cools by emission of bremsstrahlung radiation in its interior. At t = t0 , the sphere is optically thin. a) What is the total luminosity of the sphere as a function of M0 , R(t) and T0 while the sphere is optically thin? b) What is the luminosity of the sphere as a function of time after it becomes optically thick? c) Give an implicit relation, in terms of R(t), for the time Tthick when the sphere becomes optically thick. d) Draw a qualitative curve of the luminosity as a function of time. 24. A region of space contains relativistic electrons and magnetic elds. Let a typical linear scale of this region be l. Suppose the region is compressed (by passage of a shock wave, perhaps). Assume that the compression is the same in all directions. We want to see what eect this compression has on various properties of the electrons and magnetic eld. a) Show that the magnetic eld satises B l2 . b) If the compression is slow, show that momentum of an electron satises P l1 and that the magnetic ux through electron orbits is approximately conserved. c) Show that the synchrotron emission P l6 , that the critical frequency c l4 and that the half-life for the electron t1/2 l5 . (This shows that moderate compression can profoundly eect the observed emission!) 25. Compare the relationship between the pressure and energy of an ideal monatomic, nonrelativistic gas with that of a radiation eld. Do the same for the relationship between pressure and temperature during adiabatic expansion. 26. The interstellar medium of our Galaxy contains dust particles (assume they have a radius a) and starlight (assume Tc 104 K) that bathes the dust particles in dilute radiation eld. Assume that the fraction of sky covered with stellar photospheres is w = 1014 , so that J wB (Tc ). a) Estimate the equilibrium temperature of the dust particles. b) At what wavelength would a blackbody emitting at this temperature have the largest intensity? 27. Venus has a temperature of about 600 K. Radio waves with wavelengths > 30 cm can propagate easily through the atmosphere of Venus, but are absorbed and emitted by the soil at the surface of the planet. a) At approximately what frequency does the atmosphere of Venus become optically thin to radio waves? b) Remembering that F = I S , compute the radio ux at 30 cm received at Earth from Venus. c) Express your result in radio astronomy units of Janskys. (1 Jy = 1026 Watts m2 Hz1 .)

141

28.

What are the units for the Poynting Vector?

29.

What is b, the Doppler parameter, of a Doppler broadened prole for a Lyman- line emitted by a hydrogen gas with T = 104 K? a) What is , the velocity dispersion, for this gas? b) What is the FWHM (full width at half-maximum) of the prole in A? c) How does the FWHM compare to line widths determined for /2 , the natural line width for the Lyman- line? d) What values would b, , and FWHM have for the Balmer- line? e) How would these quantities change for carbon atoms ( = 1335 A)?

30.

Two important transitions of atomic hydrogen in the cosmos are the Lyman- transition (from ground state to the rst excited electronic state with = 1215 A) and the 21cm transition (a hyperne transition in which the spin of the electron changes with respect to the proton spin). a) In a stellar photosphere where the temperature is T = 6000 K, how large is the correction factor for stimulated emission of Lyman- photons? b) In an interstellar gas cloud with typical temperature is T = 100 K, how large is the correction factor for stimulated emission of 21cm photons? c) What is the ratio of absorption coecients for Lyman- and 21cm emission of hydrogen gas for temperatures below 6000 K? In this temperature regime, how does the ratio of optical depths at these two wavelengths depend on temperature, i.e., what is the functional dependence on T ?

31.

The goal of this problem is to determine the behaviour of the brightness temperature TB ( ) observed from a dust layer under dierent assumptions for the density distribution of the dust. The layer has thickness s0 and the dust has a uniform temperature T through this layer. The dust particles have number density n0 . Each dust grain has an absorption cross section that depends on frequency so that = (/0 ) 0 , where 0 is the cross section at the frequency 0 . (This type of behaviour occurs for dust when the wavelength of the radiation is larger than the size of the grain.) a) Compute the brightness temperature TB ( ) that an observer would measure as a function of frequency, and plot it. (Assume that h kT .) b) Now suppose that this absorbing layer of dust is actually a surface layer in front of a very deep and dusty cloud that has temperature T2 . Compute and sketch TB ( ) for the two cases T > T2 and T < T2 . c) Would your answers to a) and b) be dierent if the number density of the layer was not constant, but decreased with depth linearly so that n(s) = n0 (1 s/s0 )? (Layer is densest furthest away from observer.) If so, sketch how TB ( ) would be dierent.

142

32.

In this exercise, you compute the functional form E ( ), the Fourier transform of electromagnetic pulses of varies types, in order to understand corresponding frequency content of the radiation. a) What is the functional form of E ( ) for a Gaussian pulse E (t) = E0 exp (t/t)2 ? b) What is the functional form of E ( ) for a truncated wavetrain that is a constant amplitude cosine wave of (angular) frequency throughout its duration and has a total length t = N ? What is |E ( )|2 ? Show that the received radiation has a characteristic frequency width equal to 1/t, where t = 2N/0 . c) Derive the functional form of E ( ) for a damped sinusoid E (t) = E0 exp (t/t) sin 0 t for t > 0 assuming that E (t) = 0 for t < 0. What is |E ( )|2 ? What is the form of the characteristic frequency width ?

33.

Consider electron scattering in the cosmic environments described below. a) The mean electron density in the plane of the Milky Way is ne 0.01cm3 . What is the optical depth due to electron scattering in visible light along a path 10 kpc long through the plane of the Galaxy? b) What is for lines of sight through the Orion Nebula, which has a diameter of about 10 pc and ne 103 cm3 ? c) What is through a cluster of galaxies 1 Mpc in diameter containing an intergalactic medium with ne 104 cm3 ? d) A neutron star in a close binary system accretes gas from its companion star. Within about 100 km of the neutron star, the electron density is ne 1018 cm3 . What is the corresponding ? e) In which of these astrophysical environments would you expect the eects of electron scattering to be the largest? f) Why can proton scattering be ignored?

34.

Download the spectral data for your personal quasar from the Astrofysica webpage. a) Find an emission line, and measure its width. b) How large must the Doppler velocities be in order to explain the width of this spectral emission line? c) How do these speeds compare to the rotation speeds of typical spiral galaxies?

143

35.

For the Lyman- line (hydrogen atom transitions from rst excited state n = 2 to the ground state n = 1, the A21 coecient has the value 6.265 108 sec1 . a)What is the average lifetime of atoms in the n = 2 level? (Assume there is negligible background J so you can ignore stimulated emission.) b) What is the natural line width (in frequency of photons emitted by n = 2 to 1 transitions? c) What is the line width in Angstroms A, given that the L line has = 1215.6 A? Is this value greatly dierent from the classical line width? FHB 4.3

36.

For an HII Region like the Orion Nebula, with size L 10 pc, electron density ne 103 cm3 , and electron temperature T 104 K, a) What is the value of ? b) What is the total luminosity of the HI region due to free-free emission? (Assume that it is optically thin. You will check this assumption by doing the steps below.) How many solar luminosities (1L = 41033 erg/sec) does this correspond to? c) At what frequency does = 1? d) Sketch the emitted spectrum. e) Estimate the cooling time. [Hot gas can cool by emitting thermal bremsstrahlung. Assume that the only energy available is the internal thermal kinetic energy of the hot 3 gas, U = 2 nkT .] Is this a long or short time compared to astronomical time scales? What is the implication for HII regions?

37.

For a pulsar whose pulses are observed to be delayed by 1.13 sec between 400 and 300 MHz when they arrive at an earth-based radio telescope, a) What is the the value of the dispersion measure DM? b) If this pulsar is known to be 1800 pc from Earth, what is the average electron density ne along the line of sight to the pulsar?

144

38.

What is the radiation pattern dP d ( ) for cyclotron radiation, where is the angle relative to the B eld direction? Describe (qualitatively!) the polarization of the radiation emitted by the cyclotron process at dierent angles .

39. What is the amplitude of the aberration of starlight eect that is experienced by observers on Earth? How might it be detected? How does the eect dier from, say, trigonometric parallax (used for determining distances to nearby stars)? 40. This exercise is about the gyration of cosmic particles in magnetic elds. a) What is the gyrofrequency for an interstellar electron (assume B = 105 gauss)? b) For an interstellar proton? c) How big is the radius of gyration for these particles if they have thermal velocities (with T = 500K)? d) For relativistic electrons with = 1, estimate the gyrofrequency and radius of gyration in the same B = 105 gauss eld. How would you answer dierently if = 10? 41. Consider the compact radio source whose radio spectrum is sketched below. The source is 300 Mpc from Earth. The radio observations show variability on time scales of 1 month.

a) Estimate the physical size of the source. b) What is the solid angle subtended by the source? c) What is the maximum intensity? d) What is the maximum brightness temperature? e) Estimate the magnetic eld strength. f) What is the maximum particle energy? g) Explain the shape of the spectrum at < o . h) Explain the shape of the spectrum at o < < 1 . i) Explain the shape of the spectrum at > 1 . j) Estimate the total energy in particles and in the eld. (Take the ratio of energy in positive particles to electrons to be 100; do not assume equipartition.)

145

42.

In astronomical spectroscopy, we can only measure velocities of atoms along the line of sight. 2 To determine the temperature of a gas, whose mean squared random velocity < vr > along the line of sight can be determine from the shape of a spectral line, we need to know how 2 2 < vr > is related to T . Show that for a Maxwell-Boltzmann distribution, < vr >= kT /m. In an ionized hydrogen (HII) region, protons and electrons are dissociated. If the temperature of this interstellar gas is 104 K, what are the typical electron and proton velocities? Suppose that a 104 cm radius dust grain absorbs 1/3 of the solar radiation incident on its surface, and scatters the remainder isotropically. a) Calculate the ratio of gravitational attractive force to radiative repulsive force from the Sun, assuming that the grain has a density of 6 g cm3 . b) Show that this ratio is independent of the distance from the Sun. c) What is the orbital period of such a grain moving in the Earths orbit, and how does it compare to the orbital period of the Earth?

43.

44.

45.

Suppose that the magnetic eld in a region of space increases from 106 to 105 gauss over a period of 107 years. a) To what energies would electrons and protons be accelerated, if they move perpendicular to the eld and suer no collisions? [Hint: First show that the energy-eld relationship is dE/E = dB/B ] b) How does the nal energy depend on the initial energy?

46.

Show that the synchrotron half life time in the synchrotron-self-Compton model is t1/2 = 5 108 B sin2 1G
2

seconds.

(10.16)

47.

Estimate how far you could see through the Earths atmosphere if it had the opacity of the 1 2 solar photosphere ( = 0.264g cm ). 5000A An object emits a blob of material at speed v at an angle to the line of sight. a) Show that the apparent transverse velocity inferred by the observer is vapp = v sin . 1 v c cos (10.17)

48.

b) Show that vapp can exceed c and derive the angle for which vapp is a maximum.

146

49. A radio source with ux density S = 1 Jy is measured with VLBI to have an angular diameter 5 milliarcseconds at 1 GHz. What is the brightness temperature of this source? 50. Consider a source moving at velocity v at an angle to the line of sight. Quantities in the rest frame of the source are denoted with primed co-ordinates. a) Show that the frequency of radiation emitted in the rest frame of the source, , is related to the frequency in the rest frame of an observer, , by = D , D= 1 , (1 cos ) (10.18)

where D is the called the Doppler factor. Hence show that the brightness temperature in the rest frame of the observer is
TB = D TB .

(10.19)

[Hint: use the denition TB = c2 I /2 2 k and the fact that I / 3 is a Lorentz invariant.] b) Suppose a source is observed to exhibit intrinsic variability. Flux density variations of amplitude S occur on a timescale tvar . Then we can infer the size of the emitting region from variability arguments. Argue that if the source size is R, then one has tvar R/cD. Hence show that the observed brightness temperature of a source whose angular diameter is deduced from the timescale of intrinsic variability is
TB D 3 TB .

(10.20)

Show that the observed brightness temperature is Tvar = 2 S D2 , 2kc2 t2 var (10.21)

where D is the distance to the source. [NB: neglect cosmological (redshift-dependent) eects in this problem, since these modify the results only by terms of order unity.] c) The source J1819+3845 exhibits 100 mJy uctuations on a timescale of 30 minutes at a frequency of 5 GHz. If these variations were intrinsic, what would the observed brightness temperature of the source be? What Doppler factor would be required to explain this brightness temperature supposing that the source does not exceed the inverse Compton limit? d) In fact, the variability in J1819+3845 is due to interstellar scintillation, and is not intrinsic to the source. However, to exhibit scintillation, the source must have an angular diameter S 50 as, and 200 mJy of the source ux density must be coming from this compact region. What Doppler factor is now required to explain the brightness temperature of the source supposing that the emission does not exceed the inverse Compton limit? 51. Electron synchrotron radiation is limited to a brightness temperature of T 1012 K by inB verse Compton scattering. What is the corresponding limit for proton synchrotron radiation, assuming no electrons are present in the medium to re-absorb this radiation? 147

52. Show that there is no electric dipole emission associated with bremsstrahlung due to electronelectron encounters. 53. Consider the electromagnetic eld at a point described by polar co-ordinates r, due to a charge q in constant rectilinear motion at velocity c along the polar axis. Assume that the particle is at the origin r = 0 at time t = 0. (a) Show that the retarded time is given by the smaller of the solutions of the quadratic equation (t tr )2 = r2 /c2 2tr (r/c) cos + 2 t2 r. (10.22)

(b) Show that the electric and magnetic elds of a charge q in constant rectilinear motion at velocity c are given by E B with = (1 2 )1/2 . (c) In the non-relativistic limit (10.23) gives the Coloumb eld. Give a physical interpretation of the non-relativistic limit of (10.24) in terms of the Biot-Savart law. 54. Using PCompt = 4 T c 2 2 Uph . 3 (10.25) = = x r q 2 (r x )3 q r x , c 2 (r x )3 (10.23) (10.24)

show that the total power scattered by the inverse Compton process is P = 4 3a 3a T cWph K (3 a)1 max min , 3 (10.26)

for a distribution of electrons N ( ) = K a , 0, min < < max . otherwise (10.27)

148

55. Show that the transformation of acceleration is ax ay az = = = a x , 3 3 a y 2 2 a z 2 2

u yv c2 u zv c2

a x , 2 3 a x 2 3

(10.28) (10.29)

where 1+ vu x . c2 (10.30)

If K is the instantaneous rest frame of the particle, show that a a = = 3a a


2

(10.31) (10.32) (10.33)

where a and a are the components parallel and perpendicular to the direction of v respectively. 56. Here we show that the resonance condition k v s = 0 , (10.34)

for synchrotron radiation arises in a quantum mechanical treatment as a result of conservation of momentum and energy. The energy eigenvalues of a relativistic particle (ignoring spin) are = n (p ) = (m2 c4 + p2 c2 + 2n|q |B hc2 )1/2 . (10.35)

Perpendicular momentum is not conserved during emission, but both momentum parallel to the eld lines and energy are conserved. Show that this implies p = p hk, h. = n (p ) (10.36)

Now write n = n s and assume the changes in p , and n are all small. Then show that the expansion 2 + ), s + 1D h = (1 D 2 s reproduces (10.34) to lowest order in h . s = h D s +k v p p (10.37)

149

57.

Show that the reversal of the sense of polarization as a synchrotron source becomes selfabsorbed occurs at a frequency determined approximately by e
L

2 3a + 5

(10.38)

58.

Show that synchrotron absorption in vacuo cannot be negative. Negative absorption is only possible for d(2 , )/d < 0. Partially integrate
,

(, ) =

(2 )3 c2 2 3

d cos ()
1 0

d 2

(, )
,

d N () d 2

(10.39)

and show that negative absorption is possible only for d(2 R = /c = 2s/3 3 sin3 , = /mc2 and the relations

)/d < 0. Use the relations (10.40) (10.41)

K 1 (z ) + K +1 (z ) = 2K (z ), 2 K 1 (z ) K +1 (z ) = Knu (z ). z

Use the following and K (R) > 0 to complete the proof: d(2
,

/d) = 2R

3 1 2 K5/3 (R) + 2 K1/3 (R) 2 3R K2/3 (R)

(10.42)

59.

Show that the average rate of energy loss by an electron in an isotropic distribution (i.e. averaged over pitch angle) is = b2 , b= 4 e 2 2 e . 5 9 m2 ec (10.43)

The evolution of the energy spectrum N (, t) for electrons in a synchrotron source is modelled by an equation of the form N (, t) = [N (, t)] e N (, t) + Q(, t), t (10.44)

where Q(, t) is a source term and e N (, t) represents a loss term due to escape of electrons from the source. Solve (10.44) in the stationary case Q() .

150

You might also like