Você está na página 1de 210

Theoretical Astrophysics

Matthias Bartelmann
Institut fur Theoretische Astrophysik
Universitat Heidelberg

Contents
1

Macroscopic Radiation Quantities, Emission and Absorption

1.1

Specific Intensity . . . . . . . . . . . . . . . . . . . .

1.2

Relativistic Invariant . . . . . . . . . . . . . . . . . .

1.2.1

Lorentz Transformation of I . . . . . . . . . .

1.2.2

Example: The CMB Dipole . . . . . . . . . .

Einstein coefficients and the Planck spectrum . . . . .

1.3.1

Transition Balance . . . . . . . . . . . . . . .

1.3.2

Example: The CMB Spectrum . . . . . . . . .

1.4

Absorption and Emission . . . . . . . . . . . . . . . .

1.5

Radiation Transport in a Simple Case . . . . . . . . .

1.6

Emission and Absorption in the Continuum Case . . .

10

1.3

Scattering

13

2.1

Maxwells Equations and Units . . . . . . . . . . . . .

13

2.2

Radiation of a Moving Charge . . . . . . . . . . . . .

14

2.3

Scattering off Free Electrons . . . . . . . . . . . . . .

15

2.3.1

Polarised Thomson Cross Section . . . . . . .

15

2.3.2

Unpolarised Thomson Cross Section . . . . . .

16

2.4

Scattering off Bound Charges . . . . . . . . . . . . . .

17

2.5

Radiation Drag . . . . . . . . . . . . . . . . . . . . .

19

2.5.1

Time-Averaged Damping Force . . . . . . . .

19

2.5.2

Energy Transfer to a Radiation Field . . . . . .

20

Compton Scattering . . . . . . . . . . . . . . . . . . .

21

2.6

CONTENTS

2.7
3

2.6.1

Energy-Momentum Conservation . . . . . . .

21

2.6.2

Energy Balance . . . . . . . . . . . . . . . . .

22

The Kompaneets Equation . . . . . . . . . . . . . . .

24

Radiation Transport and Bremsstrahlung

27

3.1

Radiation Transport Equations . . . . . . . . . . . . .

27

3.2

Local Thermodynamical Equilibrium . . . . . . . . . .

29

3.3

Scattering . . . . . . . . . . . . . . . . . . . . . . . .

32

3.4

Bremsstrahlung . . . . . . . . . . . . . . . . . . . . .

34

3.4.1

Spectrum of a Moving Charge . . . . . . . . .

34

3.4.2

Hyperbolic Orbits . . . . . . . . . . . . . . . .

35

3.4.3

Integration over the Electron Distribution . . .

37

Synchrotron Radiation, Ionisation and Recombination

40

4.1

Synchrotron Radiation . . . . . . . . . . . . . . . . .

40

4.1.1

Electron Gyrating in a Magnetic Field . . . . .

40

4.1.2

Beaming and Retardation . . . . . . . . . . . .

41

4.1.3

Synchrotron Spectrum . . . . . . . . . . . . .

43

Photo-Ionisation . . . . . . . . . . . . . . . . . . . . .

44

4.2.1

Transition Amplitude . . . . . . . . . . . . . .

44

4.2.2

Transition Probability . . . . . . . . . . . . .

46

4.2.3

Transition Matrix Element . . . . . . . . . . .

48

4.2.4

Cross Section . . . . . . . . . . . . . . . . . .

50

4.2

Spectra

52

5.1

Natural Width of Spectral Lines . . . . . . . . . . . .

52

5.2

Cross Sections and Oscillator Strengths . . . . . . . .

52

5.2.1

Transition Probabilities . . . . . . . . . . . . .

53

5.3

Collisional Broadening of Spectral Lines . . . . . . . .

55

5.4

Velocity Broadening of Spectral Lines . . . . . . . . .

56

5.5

The Voigt Profile . . . . . . . . . . . . . . . . . . . .

57

CONTENTS
5.6
6

Equivalent Widths and Curves-of-Growth . . . . . . .

61

6.1

Boltzmann Equation and Energy-Momentum Tensor .

61

6.1.1

Boltzmann Equation . . . . . . . . . . . . . .

61

6.1.2

Moments; Continuity Equation . . . . . . . . .

62

6.1.3

Energy-Momentum Tensor . . . . . . . . . . .

64

The Tensor Virial Theorem . . . . . . . . . . . . . . .

68

6.2.1

A Corollary . . . . . . . . . . . . . . . . . . .

68

6.2.2

Second Moment of the Mass Distribution . . .

68

Ideal and Viscous Fluids

71

7.1

Ideal Fluids . . . . . . . . . . . . . . . . . . . . . . .

71

7.1.1

Energy-Momentum Tensor . . . . . . . . . . .

71

7.1.2

Equations of Motion . . . . . . . . . . . . . .

73

7.1.3

Entropy . . . . . . . . . . . . . . . . . . . . .

75

Viscous Fluids . . . . . . . . . . . . . . . . . . . . . .

76

7.2

7.2.1

7.3

58

Energy-Momentum Tensor and Equations of Motion

6.2

Stress-Energy Tensor; Viscosity and Heat Conductivity . . . . . . . . . . . . . . . . . . . .

76

7.2.2

Estimates for Heat Conductivity and Viscosity

78

7.2.3

Equations of Motion for Viscous Fluids . . . .

80

7.2.4

Entropy . . . . . . . . . . . . . . . . . . . . .

81

Generalisations . . . . . . . . . . . . . . . . . . . . .

82

7.3.1

Additional External Forces; Gravity . . . . . .

82

7.3.2

Example: Cloud in Pressure Equilibrium . . .

83

7.3.3

Example: Self-Gravitating Gas Sphere . . . . .

84

Flows of Ideal and Viscous Fluids

86

8.1

Flows of Ideal Fluids . . . . . . . . . . . . . . . . . .

86

8.1.1

Vorticity and Kelvins Circulation Theorem . .

86

8.1.2

Bernoullis Constant . . . . . . . . . . . . . .

88

8.1.3

Hydrostatic Equlibrium . . . . . . . . . . . . .

89

CONTENTS
8.1.4
8.2

8.3

8.4

4
Curl-Free and Incompressible Flows . . . . . .

90

Flows of Viscous Fluids . . . . . . . . . . . . . . . . .

91

8.2.1

Vorticity; Incompressible Flows . . . . . . . .

91

8.2.2

The Reynolds Number . . . . . . . . . . . . .

92

Sound Waves in Ideal Fluids . . . . . . . . . . . . . .

93

8.3.1

Linear Perturbations . . . . . . . . . . . . . .

93

8.3.2

Sound Speed . . . . . . . . . . . . . . . . . .

95

Supersonic Flows . . . . . . . . . . . . . . . . . . . .

96

8.4.1

Machs Cone; the Laval Nozzle . . . . . . . .

96

8.4.2

Spherical Accretion . . . . . . . . . . . . . . .

97

Shock Waves and the Sedov Solution


9.1

9.2

9.3

102

Steepening of Sound Waves . . . . . . . . . . . . . . . 102


9.1.1

Formation of a Discontinuity . . . . . . . . . . 102

9.1.2

Specific Example . . . . . . . . . . . . . . . . 104

Shock Waves . . . . . . . . . . . . . . . . . . . . . . 107


9.2.1

The Shock Jump Conditions . . . . . . . . . . 107

9.2.2

Propagation of a One-Dimensional Shock Front 108

9.2.3

The Width of a Shock . . . . . . . . . . . . . 111

The Sedov Solution . . . . . . . . . . . . . . . . . . . 112


9.3.1

Dimensional Analysis . . . . . . . . . . . . . 112

9.3.2

Similarity Solution . . . . . . . . . . . . . . . 114

10 Instabilities, Convection, Heat Conduction, Turbulence

118

10.1 Rayleigh-Taylor Instability . . . . . . . . . . . . . . . 118


10.2 Kelvin-Helmholtz Instability . . . . . . . . . . . . . . 121
10.3 Thermal Instability . . . . . . . . . . . . . . . . . . . 123
10.4 Heat Conduction and Convection . . . . . . . . . . . . 127
10.4.1 Heat conduction . . . . . . . . . . . . . . . . 127
10.4.2 Convection . . . . . . . . . . . . . . . . . . . 129
10.5 Turbulence . . . . . . . . . . . . . . . . . . . . . . . 130

CONTENTS
11 Collision-Less Plasmas

5
133

11.1 Basic Concepts . . . . . . . . . . . . . . . . . . . . . 133


11.1.1 Shielding; the Debye length . . . . . . . . . . 133
11.1.2 The plasma frequency . . . . . . . . . . . . . 134
11.2 The Dielectric Tensor . . . . . . . . . . . . . . . . . . 135
11.2.1 Polarisation and dielectric displacement . . . . 135
11.2.2 Structure of the dielectric tensor . . . . . . . . 136
11.3 Dispersion Relations . . . . . . . . . . . . . . . . . . 137
11.3.1 General form of the dispersion relations . . . . 137
11.3.2 Transversal and longitudinal waves . . . . . . 138
11.4 Longitudinal Waves . . . . . . . . . . . . . . . . . . . 139
11.4.1 The longitudinal dielectricity . . . . . . . . . . 139
11.4.2 Landau Damping . . . . . . . . . . . . . . . . 141
11.5 Waves in a Thermal Plasma . . . . . . . . . . . . . . . 142
11.5.1 Longitudinal and transversal dielectricities . . 142
11.5.2 Dispersion Measure and Damping . . . . . . . 145
12 Magneto-Hydrodynamics

147

12.1 The Magneto-Hydrodynamic Equations . . . . . . . . 147


12.1.1 Assumptions . . . . . . . . . . . . . . . . . . 147
12.1.2 The induction equation . . . . . . . . . . . . . 148
12.1.3 Eulers equation . . . . . . . . . . . . . . . . 149
12.1.4 Energy and entropy . . . . . . . . . . . . . . . 151
12.1.5 Magnetic advection and diffusion . . . . . . . 152
12.2 Generation of Magnetic Fields . . . . . . . . . . . . . 153
12.3 Ambipolar Diffusion . . . . . . . . . . . . . . . . . . 155
12.3.1 Scattering cross section . . . . . . . . . . . . . 155
12.3.2 Friction force; diffusion coefficient . . . . . . . 157
13 Waves in Magnetised Plasmas

159

13.1 Waves in magnetised cold plasmas . . . . . . . . . . . 159

CONTENTS

13.1.1 The dielectric tensor . . . . . . . . . . . . . . 159


13.1.2 Contribution by ions . . . . . . . . . . . . . . 161
13.1.3 General dispersion relation . . . . . . . . . . . 162
13.1.4 Wave propagation parallel to the magnetic field

163

13.1.5 Faraday rotation . . . . . . . . . . . . . . . . 164


13.1.6 Wave propagation perpendicular to the magnetic field . . . . . . . . . . . . . . . . . . . . 165
13.2 Hydro-Magnetic Waves . . . . . . . . . . . . . . . . . 166
13.2.1 Linearised perturbation equations . . . . . . . 166
13.2.2 Alfven waves . . . . . . . . . . . . . . . . . . 168
13.2.3 Slow and fast hydro-magnetic waves . . . . . . 169
14 Jeans Equations and Jeans Theorem

171

14.1 Collision-less motion in a gravitational field . . . . . . 171


14.1.1 Motion in a gravitational field . . . . . . . . . 171
14.1.2 The relaxation time scale . . . . . . . . . . . . 172
14.2 The Jeans Equations . . . . . . . . . . . . . . . . . . 175
14.2.1 Moments of Boltzmanns equation . . . . . . . 175
14.2.2 Jeans equations in cylindrical and spherical coordinates . . . . . . . . . . . . . . . . . . . . 177
14.2.3 Application: the mass of a galaxy . . . . . . . 178
14.3 The Virial Equations . . . . . . . . . . . . . . . . . . 179
14.3.1 The tensor of potential energy . . . . . . . . . 179
14.3.2 The tensor virial theorem . . . . . . . . . . . . 180
14.4 The Jeans Theorem . . . . . . . . . . . . . . . . . . . 182
15 Equilibrium, Stability and Disks

184

15.1 The Isothermal Sphere . . . . . . . . . . . . . . . . . 184


15.1.1 Phase-space distribution function . . . . . . . 184
15.1.2 Isothermality . . . . . . . . . . . . . . . . . . 185
15.1.3 Singular and non-singular solutions . . . . . . 186
15.2 Equilibrium and Relaxation . . . . . . . . . . . . . . . 187

CONTENTS

15.3 Stability . . . . . . . . . . . . . . . . . . . . . . . . . 188


15.3.1 Linear analysis and the Jeans swindle . . . . . 188
15.3.2 Jeans length and Jeans mass . . . . . . . . . . 190
15.4 The rigidly rotating disk . . . . . . . . . . . . . . . . 191
15.4.1 Equations for the two-dimensional system . . . 191
15.4.2 Analysis of perturbations . . . . . . . . . . . . 192
15.4.3 Toomres criterion . . . . . . . . . . . . . . . 193
16 Dynamical Friction, Fokker-Planck Approximation

195

16.1 Dynamical Friction . . . . . . . . . . . . . . . . . . . 195


16.1.1 Deflection of point masses . . . . . . . . . . . 195
16.1.2 Velocity changes . . . . . . . . . . . . . . . . 197
16.1.3 Chandrasekhars formula . . . . . . . . . . . . 198
16.2 Fokker-Planck Approximation . . . . . . . . . . . . . 200
16.2.1 The master equation . . . . . . . . . . . . . . 200
16.2.2 The Fokker-Planck equation . . . . . . . . . . 201

Chapter 1
Macroscopic Radiation
Quantities, Emission and
Absorption
1.1

Specific Intensity

to begin with, radiation is considered as a stream of particles;


energy, momentum and so on of this stream will be investigated
as well as changes of its properties;
a screen of area dA is set up; which energy is streaming per time
interval dt enclosing the angle with the direction normal to the
screen into the solid angle element d and within the frequency
interval d?
we begin with the occupation number: let n~p be the number density of photons with momentum ~p and the polarisation state
( = 1, 2);

further reading: Shu, The


Physics of Astrophysics, Vol I:
Radiation, chapter 1; Rybicki,
Lightman, Radiative Processes
in Astrophysics, chapter 1; Padmanabhan, Theoretical Astrophysics, Vol. I: Astrophysical
Processes, sections 6.16.3

the occupation number is the


number density of occupied
states per phase space element

the energy per photon is E = h = cp (because the photon has


zero rest mass); thus
p = |~p| =

h
;
c

(1.1)

the volume element in momentum space is d3 p; the number of


independent phase-space cells is
d3 p
p2 dpd 2 dd
=
=
,
(2~)3
h3
c3

(1.2)

where momentum has been expressed by frequency in the last


step;

Heisenbergs uncertainty relation implies that points in phase


space cannot be observed; rather,
observable cells in phase space
have a finite volume

CHAPTER 1. MACROSCOPIC RADIATION QUANTITIES, EMISSION AND ABSORPTION2


in terms of these quantities, the following amount of energy is
flowing through the screen: (number of phase space cells) times
(photon occupation number) times (energy per photon) times
(volume filled by the photons); thus
2
2 dd X
n~p h dA cos dt
dE =
c3
=1

(1.3)

the energy flowing through the screen per unit time, frequency
and solid angle is
2

X
dE
h3
=
n~p 2 cos I cos ,
dtddAd =1
c

(1.4)

where I is called specific intensity of the radiation;


for unpolarised light, we obviously have
I =

1.2
1.2.1

2h3
n~p ;
c2

(1.5)

Relativistic Invariant
Lorentz Transformation of I

switching from one reference frame to another, the transformation


properties of the physical quantities is important to be known; we
shall now show by Lorentz transformation that the quantity
I
3

(1.6)

is relativistically invariant;
let us assume two observers O and O0 , which are moving relatively to each other with velocity v in x3 direction; O0 is collecting
photons on a screen dA0 in the x10 -x20 plane which move under the
angle 0 with respect to the area normal into the solid angle d0 ;
he finds
p02 dp0 d0 0
dN 0 = 2
n p0 dA0 c cos 0 dt0
(1.7)
(2~)3
photons on his screen;
likewise, observer O expects the same screen to collect the photon
number
p2 dpd
dN = 2
n p dA(c cos v) dt0
(1.8)
(2~)3
and of course the two numbers must be equal, dN 0 = dN;

CHAPTER 1. MACROSCOPIC RADIATION QUANTITIES, EMISSION AND ABSORPTION3


the Lorentz transformation relating O and O0 is

0 0
0 1 0 0

0
0
1
0

0 0

(1.9)

with v/c and (1 2 )1/2 ;


for the screen at rest in O0 , dx30 = 0, thus
dx0 = cdt = dx00 = cdt0 ,

(1.10)

so that dt = dt0 , which is the usual relativistic time dilation;


energy and momentum are combined in the four-vector
E  

E

, ~p p0 , ~p ; p0 = |~p| because |~p| = , (1.11)


p =
c
c
and we obtain
p0 = (p00 + p03 ) ;

p3 = (p00 + p03 ) ,

(1.12)

and the other components are p1 = p01 , p2 = p02 ;


since, from simple geometry, p03 = cos 0 |~p0 | = p00 cos 0 and
p3 = p0 cos , we then find
cos =

p3 p00 + p03
+ cos 0
=
=
p0
p00 + p03 1 + cos 0

(1.13)

for the Lorentz transformation of the angle ;


this implies for the solid-angle element
#
"
d2 0
+ cos 0
2
0
=
d = d(cos )d = d d
;
1 + cos 0
2 (1 + cos 0 )2
(1.14)
summarising, we find for the number of photons in the system O:
2 
d2 0
0 3 02
0
(1
+

cos

)
p
dp
{z
} 2 (1 + cos 0 )2
h3 |
|
{z
}
=p2 dp
=d2
! #
"
+ cos 0
0
v dt0 ;
n p |{z}
dA c
(1.15)
|{z}
1 + cos 0
=dA |
{z
} =dt

dN =

=c cos v

equating this to dN from (1.7) yields


!#
"
+ cos 0
0
0
2
0
n p0 cos = (1 + cos )
np
1 + cos 0
= 2 (1 2 ) cos 0 n p = n p cos 0 ;
(1.16)

CHAPTER 1. MACROSCOPIC RADIATION QUANTITIES, EMISSION AND ABSORPTION4


thus, the occupation number is obviously a relativistic invariant,
n p = n0p0 ,

(1.17)

and I 3 n p implies the claimed invariance (1.6),


I00
I
=
;
(1.18)
3 03
the Lorentz transformation of the solid angle (1.14) will be used
later in the discussion of synchrotron radiation

1.2.2

Example: The CMB Dipole

this relativistic invariance of I /3 allows the dipole of the cosmic


microwave background to be computed: a photon flying at an
angle relative to the x axis of the observer will be redshifted by
an amount which directly follows from Lorentz transformation;
using p = (E/c, ~p) and p1 = |~p| cos , the Lorentz transformation
yields

0 0 E/c
0 0 p1

p0 =
2

0
0
1
0
p

3
0 0 0 1
p

E/c + |~p| cos


E/c + |~p| cos
,
=
(1.19)

p2

p3
i.e. the energy in the primed system is
 E

E
E 0 = c + cos = (1 + cos )E ;
c
c
the frequency is thus increased to
0 = (1 + cos ) ;

(1.20)

(1.21)

with the occupation number n p being a relativistic invariant,


1
1
n p = h/kT
= h0 /kT 0
= n0p0 ,
(1.22)
e
1 e
1
the temperature T must change exactly as the frequency , thus
T 0 = T (1 + cos ) ;

(1.23)

for non-relativistic velocities v  c, 1, and thus




v
T 0 T 1 + cos ;
(1.24)
c
the motion of the Earth relative to the microwave background thus
causes a dipolar pattern in its measured temperature; with v
103 c and T 3 K, the amplitude of the dipole is of order a few
milli-Kelvins;

CHAPTER 1. MACROSCOPIC RADIATION QUANTITIES, EMISSION AND ABSORPTION5

1.3

1.3.1

Einstein coefficients and the Planck spectrum


Transition Balance

we consider mean transition rates in an emission- and absorption


process between two energy levels E1 and E2 ; the rates of absorption and of stimulated emission will be proportional to the specific
intensity, absorption rate I B12 and stimulated emission rate
I B21 , while the rate of spontaneous emission will not depend on
I , spontaneous emission rate A21 ; A and B are called the Einstein coefficients;
now, let N1 and N2 be the mean number of states with the energies
E1 and E2 ; equilibrium between transitions will require as many
transitions from E1 to E2 as there are from E2 to E1 , thus
N1 I B12 = N2 [A21 + I B21 ] ,

(1.25)

which can be satisfied if the specific intensity is


I =

N2 A21
=
N1 B12 N2 B21

A21
A21
,

=
N1
N1
B

1
B
12
21
21 N2
N2

(1.26)

where we have used that B12 = B21 (E1 and E2 are eigenstates of
the Hamilton operator);
according to the definition of A21 and B21 , we must have
[cf. Eq. (1.5)]
h3
A21 = 2 2 B21 ;
(1.27)
c
if there is thermal equilibrium between the states E1 and E2 , we
have the Boltzmann factor between N1 and N2 ,
!
N2
h
,
(1.28)
= exp
N1
kT
where E2 = E1 + h;
under this condition, (1.28) implies
I =

2h3
1
B ,
2
h/kT
c e
1

(1.29)

which is the Planck spectrum;


limiting cases of the Planck spectrum for high and low frequencies are
B

2h3 h/kT
e
c2

for 

kT
h

(Wiens law)

(1.30)

stimulated emission is a consequence of the Bose character


of photons: if a quantum state
is occupied by photons, an increase in the occupation number
is more likely

CHAPTER 1. MACROSCOPIC RADIATION QUANTITIES, EMISSION AND ABSORPTION6


and
B

22
kT
c2

for 

kT
h

(Rayleigh-Jeans law)

(1.31)

the spectral energy density is


dU =

dE
dE
I 2
=
=
d ,
ddx3 ddA(cdt)
c

thus
cU =

I d2 ,

(1.32)

(1.33)

which equals 4I for isotropic radiation;


a unit for the spectral energy density which is frequently used in
astronomy is the Jansky, defined by
1 Jy = 1026

1.3.2

W
erg
= 1023
;
2
m Hz
cm2 s Hz

(1.34)

Example: The CMB Spectrum

for example, the spectral energy density of the CMB is given by


4
4 2h3
1
B =
c
c c2 eh/kT 1
erg
= 2.4 1025
= 23.9 mJy
cm2 s Hz
at a frequency of = 30 GHz;
U =

(1.35)

the maximum of the Planck spectrum is located at


h
2.82 ;
kT
for the CMB, this corresponds to a frequency of
x

1.60 1011 Hz = 160 GHz ;

(1.36)

(1.37)

inserting the Planck spectrum for I in (1.28) and integrating over


all frequencies yields
Z
2 (kT )4
U=
(1.38)
U d =
15 (~c)3
0
for the energy density of a Planckian radiation field;
the number density of the photons is clearly
!3
Z
2(3) kT
U
n=
d =
,
h
2
~c
0

(1.39)

where the Riemann function takes the numerical value (3)


1.202;

note that 1 Jy is not the unit of


specific intensity, which would
be Jy/sr

CHAPTER 1. MACROSCOPIC RADIATION QUANTITIES, EMISSION AND ABSORPTION7


for the cosmic microwave background, T = 2.7 K, and thus
n 400 cm3 ,

U 4.0 1013

erg
;
cm3

(1.40)

the Rayleigh-Jeans law is often used to define a radiation temperature T rad by requiring
22
!
kT rad = I ;
2
c

(1.41)

obviously, this agrees well with the thermodynamic temperature


if h/kT  2.82 and I = B , but the deviation becomes considerable for higher frequencies;

1.4

Absorption and Emission

the absorption coefficient is defined in terms of the energy absorbed per unit volume, time and frequency from the solid angle
d2 ,
!
dE
;
(1.42)
I = 3
d xdtdd2 abs
since the stimulated emission is also proportional to I , an analogous definition applies for the induced emission,
!
dE
ind
I = 3
;
(1.43)
d xdtdd2 ind
for the spontaneous emission, we define the emissivity
!
dE
,
j = 3
d xdtdd2 spn

(1.44)

i.e. the spontaneous energy emission per unit volume, time and
frequency into the solid-angle element d2 ;
the effective net absorption is
ind
net
= ;

(1.45)

energy
,
time area frequency solid angle

(1.46)

since the unit of I is

must obviously have the dimension (length)1 ; the mean free


path for a photon of frequency is thus approximately 1
;

CHAPTER 1. MACROSCOPIC RADIATION QUANTITIES, EMISSION AND ABSORPTION8


let be the cross section of an atom, molecule or other particle
for the absorption of light of frequency , then
= n ,

(1.47)

where n is the number density of absorbing systems and is their


mass density; is called opacity, whose physical meaning is
the absorption cross section per unit mass,
cm2
[] =
;
g

(1.48)

if matter is in equilibrium with a radiation field, the emitted and


absorbed amounts of energy must equal, hence
j + ind
I = I

j = net
I ;

(1.49)

using (1.28) and (1.29), we then find


!1

j
h3 N1
I = net = 2 2
1

c N2

(1.50)

i.e. if the occupation numbers are known, the emission and absorption properties in equilibrium can be calculated, and vice
versa;
in particular, in thermal equilibrium with matter, we have
I = B

1.5

net
=

j
;
B

(1.51)

Radiation Transport in a Simple Case

we consider an emitting and absorbing medium which does not


scatter for now and is being irradiated by a light bundle; let the
medium be characterised by an emissivity j and a net absorption
coefficient net
;
per unit of the traversed distance, the intensity of the light bundle
changes according to
dI = j dl
I dl
},
|{z} |{z
emission

(1.52)

absorption

from which we obtain the equation of radiation transport in its


simplest case,
dI
= j I .
(1.53)
dl

CHAPTER 1. MACROSCOPIC RADIATION QUANTITIES, EMISSION AND ABSORPTION9


the homogeneous equation (1.53) is easily solved:
dI
= I
dl

d ln I = dl ,

thus
I = C1 exp

(1.54)

!
dl ;

(1.55)

for solving the inhomogeneous equation (1.53), we assume C1 =


C1 (l) and find
"
!#
Z
dI
d
C1 (l) exp dl
=
(1.56)
dl
dl
!
Z
 0

= C1 (l) C1 (l) exp dl
!
Z
!
= j I = j C1 (l) exp dl ;
this implies
Z
C10 (l) exp

!
dl = j ,

which has the solution


!#
"
Z
Z
dl + C2 ;
C1 (l) =
dl j exp

(1.57)

(1.58)

if is a constant along the light path, the integral is simply


Z
dl = l ,
(1.59)
and then we have
C1 (l) =

j l
e ,

I (l) =

j
C2 e l ;

(1.60)

for example, if the intensity satisfies the boundary condition I =


0 at l = 0, the intensity as a function of path length becomes

j 
I (l) =
1 e l ;
(1.61)

interesting limiting cases: let L be the entire path length through


the medium; if
L  1

L  1

j
( L) = j L ,

j
I (L) =

I (L) =

(1.62)

the former is the optically thin, the latter the optically thick
case; this amounts to comparing the mean free path 1
to the
total path length L;

CHAPTER 1. MACROSCOPIC RADIATION QUANTITIES, EMISSION AND ABSORPTION10


if the radiation is in thermal equilibrium with the irradiated material, we must have

j 
I = B =
1 e L ,
(1.63)

which implies that the source emits at most the intensity of the
Planck spectrum
we consider optically thin, thermal emission of radio waves; optically thin implies L  1 and I = j L, thermal equilibrium
requires I = B , and in the radio regime we have
h
1,
kT

22
kT ;
c2

(1.64)

combining these conditions, we find


I j L = B L =

22
22

kT
L
=
kT b ,

c2
c2

(1.65)

where T b is the observed temperature, which is obviously related


to the emission temperature T by
T b LT ;

(1.66)

this absurd conclusion shows indicates that the two assumptions,


thermal equilibrium and optically-thin radiation, are in conflict
with each other;

1.6

Emission and Absorption in the Continuum Case

in the discrete case, the energy balance for the emitted energy was
N2 A21

h12
= E
|{z}
|{z}
transition number energy per transition

(1.67)

the emissivity (per unit solid angle) is


j =

N2 A21 h12
N2 A21 h

D ( 12 ) ,
4
4

(1.68)

with the Dirac delta function modeling a sharp line transition;


correspondingly, we generalise this expression by a line profile
function (),
N2 A21 h
j =
() ,
(1.69)
4
where () quantifies the transition probability as a function of
frequency;

CHAPTER 1. MACROSCOPIC RADIATION QUANTITIES, EMISSION AND ABSORPTION11


by an analogous procedure for the absorption coefficient, we find
=

N1 B12
h () ;
4

(1.70)

we now consider an electron of energy E which emits the energy


d
P(, E)
ddt

(1.71)

per unit time and unit frequency; let further f (~p) be the momentum distribution of the electrons, then the number of electrons
with energies between E and E + dE is
n(E)dE = f (~p)

d3 p
dp
dE = 4p2
f (~p) dE ,
dE
dE

(1.72)

if we assume the distribution to be isotropic in momentum space;


since each electron emits the energy
d = P(, E) ddt ,
we obtain for the emissivity
Z
Z
4 j =
n(E)P(, E)dE = 4
0

p2 f (p)

(1.73)

dp
P(, E)dE
dE
(1.74)

by definition, we have for a continuous transition


Z E2
P(, E2 ) = h
A21 ()dE1 ,

(1.75)

i.e. electrons with the energy E2 can emit in transitions to all possible states with E1 < E2 ; thus
Z
2h3 E2
P(, E2 ) = h 2
B21 ()dE1 ;
(1.76)
c
0
likewise, the net absorption coefficient is

Z
Z

h
=
dE1 dE2 n(E1 )B12 n(E2 )B21 () ;
| {z }
4
| {z }
absorption stim. emission
(1.77)
the second term in this expression can be written
Z
Z
h
dE1 dE2 n(E2 )B21 ()
4
Z
Z E2
h
=
dE1 B21 ()
dE2 n(E2 )
4
0
Z
c2
=
dE2 n(E2 )P(, E2 ) ,
8h3

(1.78)

CHAPTER 1. MACROSCOPIC RADIATION QUANTITIES, EMISSION AND ABSORPTION12


while the first term reads
Z
Z
h
dE1 dE2 n(E1 )B12 ()
4 Z
Z
h
=
dE2 n(E2 h) dE1 B12 ()
4
Z
c2
dE2 n(E2 h)P(, E2 ) ;
=
8h3
we thus obtain for the absorption coefficient
Z
c2
dE [n(E h) n(E)] P(, E) ;
=
8h3

(1.79)

(1.80)

in thermal equilibrium and far from the Fermi edge, the electron
number density is
 E
n(E) exp
,
(1.81)
kT
thus
"
!
#
h
1 ,
n(E h) n(E) = n(E) exp
kT

(1.82)

from which we obtain

Z
c2  h/kT
=
e
1
dE n(E) P(, E)
8h3

j
c2  h/kT
,
e

1
j =
=
3
2h
B

just as in the discrete case;

(1.83)

Chapter 2
Scattering
2.1

Maxwells Equations and Units

we use cgs units, i.e. the dielectric constant and the magnetic permeability of the vacuum are both unity, 0 = 1 = 0 ; Maxwells
equations in vacuum then read
~ E~ = 4 ,
~ B
~=0,

~
~
~ E~ = 1 B ,
~ B
~ = 4 ~j + 1 E ,

c t
c
c t

(2.1)

where is the charge density and ~j is the current density;


the Lorentz force per unit charge is
~v ~
;
f~L = E~ + B
c

(2.2)

the energy density of the electromagnetic field is


U=

1  ~ 2 ~ 2
E +B ;
8

(2.3)

~ have dimension
consequently, the field components E~ and B
!1/2 
 erg 1/2
g 1/2
g cm2
=
=
(2.4)
cm3
s2 cm3
cm s2
forces have the dimension
g cm
dyn ;
s2

(2.5)

thus, the Lorentz force


~
[F~L ] = [q][E]

 g 1/2
g cm
= [q]
,
s2
cm s2
13

(2.6)

further reading: Rybicki, Lightman, Radiative Processes in


Astrophysics, chapter 7; Padmanabhan, Theoretical Astrophysics, Vol. I: Astrophysical
Processes, sections 6.46.7

CHAPTER 2. SCATTERING

14

implies that the unit of charge must be


[q] =

g1/2 cm3/2
;
s

(2.7)

in these units, the elementary charge is


e = 4.8033 1010

cm3/2 g1/2
;
s

(2.8)

the Poynting vector, i.e. the vector of the energy current density
of the electromagnetic field, is
c  ~ ~
S~ =
EB ,
(2.9)
4
with dimension
cm erg
[S~ ] =
(2.10)
s cm2 s
which is obvious because the unit of E~ 2 is
erg
;
(2.11)
[E~ 2 ] =
cm3

2.2

Radiation of a Moving Charge

far from its source, the electric field of an accelerated charge is,
in the non-relativistic limit |~|  1


q 

~
~
~e ~e ,
E=
(2.12)
cR
where ~e is the unit vector pointing from the radiating charge to
the observer, and R is the distance;
~ = ~e E~ and E~ = B
~ ~e in vacuum, the B
~ field is
since B


~ = q ~e ~ ,
B
cR
and the Poynting vector is

i
h
i
c h ~
~ = c B
~ 2~e ( B
~ ~e) B
~ = c B
~ 2~e
S~ =
B ~e B
4
4
4
~ = 0;
because ~e B

(2.13)

(2.14)

per unit time, the energy


dE ~ ~
= S dA
(2.15)
dt
~ since dA
~ is related to the
is radiated through the area element dA;
2
~ as dA
~ = R ~ed, we find
solid-angle element d
dE
c ~2 2
=
B R d ;
(2.16)
dt
4
thus, the energy radiated per unit time into the solid-angle element
d is
c ~ 2
dE
=
(2.17)
R B ;
dtd 4

see, for example, Jackson,


Classical
Electrodynamics,
eq. (14.18)

CHAPTER 2. SCATTERING

15

2.3

Scattering off Free Electrons

2.3.1

Polarised Thomson Cross Section

a point charge q is accelerated by an incoming electromagnetic


wave with the electric field component E~ 0 ; the equation of motion
for the charge is


~ 0 qE~ 0 + O() ,
m~x = F~L = q E~ 0 + ~ B
(2.18)
i.e. the last approximation employs the non-relativistic limit of
the Lorentz force; thus, the acceleration is
q
~x = c~ = E~ 0 ;
m

(2.19)

using the dipole moment d~ q~x, we can write eq. (2.13) for the
magnetic field in the form

~e d~
~
B= 2 ;
cR

(2.20)

according to (2.19), the second time derivative of the dipole moment is


2
q
(2.21)
d~ = E~ 0 ,
m
which, when combined with (2.20) and (2.17), implies
!2
dE
c 1

~
=
~e d
dtd
4 c2
2
q4
q4 ~ 02 2
0
~
~
=
e

E
=
E sin , (2.22)


4c3 m2
4c3 m2
where was introduced as the angle between the incoming electric field E~ 0 and the direction of the outgoing radiation, ~e;
the incoming energy current density is
S0 =

c ~ 02
E ;
4

thus the differential scattering cross section is


!2
d
1 dE
q2
=
=
sin2 ;
d S 0 dtd
mc2

(2.23)

(2.24)

suppose the elementary charge e is homogeneously distributed


on the surface of a sphere with radius re ; then, its absolute potential energy is
e2

;
(2.25)
re

CHAPTER 2. SCATTERING

16

equating this to an electrons rest-mass energy me c2 , we can solve


for re ,
e2 !
= me c2
re

re =

e2
2.8 1013 cm ;
me c2

(2.26)

this is the so-called classical electron radius;


generally, the radius
q2
(2.27)
mc2
is associated with a particle of charge q and rest-mass m; using
this radius, the differential scattering cross section reads
r0

d
= r02 sin2 ;
d
the total cross section is
Z
Z
2
2
2
= r0 sin d = 2r0
0

sin3 d =

(2.28)

8 2
r ;
3 0

for electrons, we obtain the Thomson cross section,


!2
8 2 8 e2
6.6 1025 cm2 ;
r =
T =
3 e
3 me c2

2.3.2

(2.29)

(2.30)

Unpolarised Thomson Cross Section

this scattering cross section is valid for one particular polarisation


direction; we now average over all incoming polarisation directions; for doing so, we introduce the angle in the plane perpendicular to the incoming direction ~n0 ; the polarisation direction is
then

cos

~e0 = sin ,
(2.31)

0
if ~e0 is parallel to the x3 axis; the outgoing direction of the scattered radiation is

sin

~e = 0 ;
(2.32)

cos
using this, one obtains the differential scattering cross section
h
i
d
= r02 sin2 = r02 (1 cos2 ) = r02 1 (~e ~e0 )2 ;
d

(2.33)

CHAPTER 2. SCATTERING

17

using ~e ~e0 = sin cos , averaging over yields


* +
Z 2 

r02
d
=
d 1 sin2 cos2
d
2 0
"
#
Z
sin2 2
2
2
= r0 1
d cos
2 0
r02
=
(1 + cos2 ) ;
2

(2.34)

this is the unpolarised Thomson cross section;

2.4

Scattering off Bound Charges

an accelerated charge radiates energy and thus damps the incoming, accelerating wave; the non-relativistic Larmor formula
asserts that a non-relativistic, accelerated charge q emits the
power
2q2 2
P = 3 ~v ;
(2.35)
3c
this is interpreted as damping with a force F~D ,
F~D ~v = P

2q2 2
~v F~D = 3 ~v ;
3c

the temporal average over a time interval T is


* +
Z
dE
1 T 2q2 2
=
dt 3 ~v
dt
T 0
3c
"
#
2
T Z T
1 2q
~v~v 0
=
dt~v~v ;
T 3c3
0

(2.36)

(2.37)

the first term vanishes for bound charges and large T , thus
D

E 2q2 D... E
F~D ~v = 3 ~x ~v ;
3c

(2.38)

we thus identify the expression


2q2 ...
F~D = 3 ~x
3c

(2.39)

with the time-averaged damping force;


for bound orbits with an angular frequency of 0 , we have
...
~x = 20 ~x ~x = 20 ~x ;
(2.40)

cf. Jackson, Classical Electrodynamics, eq. (14.22)

CHAPTER 2. SCATTERING

18

thus, the equation of motion reads


q
~x + 20 ~x = E~ 0 eit ~x
m

(2.41)

with the damping term


=

2q2 2
;
3c3 0

(2.42)

the first term on the right-hand side of (2.41) is the external excitation, the second is the damping; this equation models a driven,
damped harmonic oscillator, whose solution is known to read
q
E~ 0 eit
~x =
;
m 20 2 i

(2.43)

we put this back into Larmors equation (2.35) and obtain


dE
2q2 2 2q2
~x = 3 ~x ~x
=P =
dt
3c3
3c
4
2q4 ~ 2
=
;
E
3m2 c3 0 (2 20 )2 + 2 2

(2.44)

the incoming energy current is |S~ | = cE~ 02 /(4), and thus the scattering cross section becomes
1 dE 8 2
4
=
=
r
3 0 (2 20 )2 + 2 2
|S~ | dt

(2.45)

with the typical resonance behaviour at = 0 ;


interesting limiting cases are:
 0 :

 0 :

0 :

8 2
r = T ;
3 0
(binding forces are then irrelevant)
!4

T
;
0
(Rayleigh scattering)
20
T
4( 0 )2 + 2
#
"
22 q2
/(2)
=
,
(2.46)
mc ( 0 )2 + (/2)2

where the term in square brackets defines the so-called Lorentz


profile;

CHAPTER 2. SCATTERING

2.5
2.5.1

19

Radiation Drag
Time-Averaged Damping Force

in the case of Thomson scattering, the scattering charge damps the


motion which is caused by the incoming electric field according
to the damping force (2.39)
2q2 ...
FD = 3 ~x ;
3c

(2.47)

an incoming electromagnetic wave exerts the Lorentz force


~ = m~x
FL = q(E~ ~ B)

(2.48)

on the scattering charge; the last equality in (2.48) assumes that


F~D  F~L , i.e. the back reaction of the radiation by the charge was
neglected;
from (2.48), we find
...
q  ~ ~ ~ ~ ~ 
~x =
E+B+B
m

q ~ ~ ~ ~x ~
E + B + B
=
m
c

q  ~ ~ ~ ~
q ~ ~ ~
E+B B ;
E+B+
=
m
mc

(2.49)

in the non-relativistic limit, we can drop the terms proportional to


~ and find
... q 
q ~ ~ 
~
~x =
E+
(E B) ,
(2.50)
m
mc
and thus
2q3  ~
q ~ ~ 
F~D =
E
+
(E B) ;
(2.51)
3mc3
mc
averaging over time yields
2
2q3
~ + 2 q
hF~D i =
h
Ei
3mc3 |{z} 3 mc2

!2

~ ;
hE~ Bi

(2.52)

=0

using
~ = hE~ + (~e E)i
~ = E~ 2~e (E~ ~e) E~
hE~ Bi
|{z}
=0

= 4U~e ,

(2.53)

we finally find
8 2
r U~e = T U~e
3 q
for the time-averaged damping force;
hF~D i =

(2.54)

CHAPTER 2. SCATTERING

2.5.2

20

Energy Transfer to a Radiation Field

we now consider a charge moving with a relative velocity ~v


through a radiation field which is isotropic in its rest frame; in
the rest frame of the radiation, we have
~ = 0 = h Bi
~ ,
hEi

~ 2i ;
hE~ 2 i = 4U = h B

(2.55)

in the rest frame of the charge, the Lorentz force is


~ 0 ) = qE~ 0 ,
F~L0 = q(E~ 0 + ~0 B

(2.56)

because ~0 = 0 in the charges rest frame; in addition, we have




~ + E~ k ,
E~ 0 = E~ 0 + E~ k0 = E~ + ~ B
(2.57)
where is the usual Lorentz factor; for Larmors equation, we
further need
F~L0
dE 2q2 0 2
0



= 3 ~x , ~x =
;
(2.58)
dt
3c
m
in a first step, we compute
q2
m2
q2
=
m2
q2
=
m2
q2
=
m2

D E
~x0 =

h
i2 
~ + E~ k
(E~ + ~ B)
h
i2 
~
~
~
~
~
(E Ek + B) + Ek
h
i2 
~ + (1 )E~ k
(E~ + ~ B)

(2.59)

h
i
~ 2 ihsin2 i + (1 2 )hE~ k2 i ,
2 hE~ 2 i + 2~2 h B

where we have used that


~ =0
hE~ k (~ B)i

(2.60)

because the direction of ~ is random with respect to the direction


~ thus, we obtain
of E~ B;
!
 2 
22 1 2
q2
~x0
2
+
= 4 U 2 1 +
m
3
32
!
q2
2
= 42 U 2 1 +
;
(2.61)
m
3
with that, we find the result
!
!
dE
2q4
2
2
2
2
= 4 U 2 3 1 +
= T Uc 1 +
dt
3m c
3
3

(2.62)

for the radiation which is on average radiated by the charge;

CHAPTER 2. SCATTERING

21

according to the radiation damping, the energy which is on average absorbed by the charge is
!
dE
= T Uc ,
(2.63)
dt abs
and thus the total energy change of the radiation field per unit
time is
!
#
"
4
2
dE
2
1 = T Uc2 2 ;
= T Uc 1 +
(2.64)
dt
3
3
this amount of energy is added to the radiation field per unit time
by a single charge;
the number of collisions between the charge and photons per unit
time is
dNc
U
= T c ;
(2.65)
dt
h
combining this with (2.64), we find the energy gain of the radiation field by scattering of the charge per scattering process,
D
E dEc dNc !1 4
4
= h2 2 = 2 2 E ;
E =
(2.66)
dt
dt
3
3

2.6
2.6.1

Compton Scattering
Energy-Momentum Conservation

we now consider electromagnetic radiation as being composed of


photons; if an ensemble of charges is embedded into a radiation
field, energy is transfered by scattering from the photons to the
charges and back; if the radiation temperature is higher than the
temperature of the charge ensemble, energy flows from the radiation to the charges; this process is called Compton scattering;
in astrophysics, the inverse Compton scattering is typically more
important, during which energy is transfered from the charges to
the radiation field;
an incoming photon with momentum h~e/c hits an electron with
momentum ~p; after scattering, the photon and the electron have
momenta h0~e0 /c and ~p0 ;
conservation of momentum and energy imply
h~e + c~p = h0~e0 + c~p0 ,

h + E = h0 + E 0 ,

(2.67)

where
E 2 = c2 p2 + m2 c4
according to the relativistic energy-momentum relation;

(2.68)

CHAPTER 2. SCATTERING

22

solving the energy equation for E 02 and inserting (2.68) yields


c2 ~p02 = c2 ~p2 + h2 ( 0 )2 + 2Eh( 0 ) ,

(2.69)

while the momentum equation implies


c2 ~p02 = c2 ~p2 + h2 (~e 0~e0 )2 + 2h(~e 0~e0 )c~p ;

(2.70)

subtracting (2.69) from (2.70) and cancelling suitable terms gives


h0 (1 cos ) = E( 0 ) c~p(~e 0~e0 ) ,

(2.71)

where is the angle between ~e and ~e0 ;


if the electron is originally at rest, ~p = 0 and E = mc2 , and (2.71)
simplifies to
h
0 (1 cos ) ,
(2.72)
0 =
2
mc
and in the limit of very low photon energy, h  mc2 , we find for
the relative energy change of the Compton-scattered photon
E 0
E
=
= 2 (1 cos ) ,
E

mc

(2.73)

and if h & mc2 , quantum electrodynamics must be used anyway;


averaging (2.73) over all scattering angles , we find the mean
energy loss per Compton scattering,
hE i =

2.6.2

E2
mc2

(1 cos )d(cos ) =

E2
mc2

(2.74)

Energy Balance

the total energy transfer to the radiation field due to the motion of
a single charge is given by the difference between the energy gain
(2.66) per scattering and the energy loss per Compton scattering
(2.74),
!
D
E
4 2 2 E
E = 2 E ;
(2.75)
3
mc
for photons with E  mc2 in the relativistic limit, 1, and
D

E 4
E 2 E ,
3

(2.76)

which can become a very large number; in that way, for example,
CMB photons can be converted to X-ray photons;

CHAPTER 2. SCATTERING

23

in the thermal limit of (2.75), we can approximate v  c, thus


1, and mv2 = 3kT e ; then
!
D
E

 E
E
4v2
E

E
=
4kT

E
;
(2.77)

3c2 mc2
mc2
thus, the photons gain energy (on average), if
4kT e > E

(2.78)

(inverse Compton scattering), and lose energy otherwise (Compton scattering)


Compton scattering causes fast charges to lose energy; typical
time scales are, according to (2.64)
tc

E
3 mc2
mc2
=
= 4
;
dE/dt
Uc2 2 4 2 T U
3 T

(2.79)

for non-relativistic, thermal electrons, E = 3kT e /2 and 1, and


tc =

3
kT e c2
2
4
Ucv2
3 T

9 mc
;
8 T U

(2.80)

after Ns scatterings, the total energy transfer from thermal electrons to the photons is
E0
4kT e
= 1+
E
mc2

!Ns

!
4kT e Ns
exp
e4y ,
mc2

(2.81)

where the Compton parameter


y

4kT e Ns
mc2

(2.82)

was introduced;
if the electron number density is ne , the number of scatterings per
path length dl is
Z
dNs = ne T dl Ns =
T ne dl ,
(2.83)
and thus the Compton-y parameter becomes
Z
kT e
y=
T ne dl ;
mc2

(2.84)

CHAPTER 2. SCATTERING

2.7

24

The Kompaneets Equation

we need an additional equation which specifies how the photon


spectrum is changed due to the scatterings; for deriving it, we
assume that a homogeneous, thermal distribution of electrons is
located in a homogeneous sea of radiation, such as, for example,
a galaxy cluster in the microwave background; the collisions with
the electrons change the photon energy, but not their number, and
thus their spectrum cannot remain a Planck spectrum;
let n() be the occupation number of photon states with frequency
; then, the Boltzmann equation requires
!
Z
Z
n()
d
3
c
(2.85)
=
d p d
t
d




n(0 ) [1 + n()] N(E 0 ) n() 1 + n(0 ) N(E) ;
this equation has the following meaning: the occupation number
at the frequency changes due to scattering from to 0 , and from
0 to ; the term


n() 1 + n(0 ) N(E)
(2.86)
quantifies how many photons there are at frequency , corrected
by the factor for stimulated emission from to 0 , and multiplies
with the number of collision partners N(E) at energy E; in other
words, it quantifies the number of collisions away from frequency
; analogously, the term
n(0 ) [1 + n()] N(E 0 )

(2.87)

quantifies the opposite scattering, i.e. scattering processes increasing the occupation number at frequency ; of course, the
energy difference between photon frequencies and 0 must be
balanced by the difference between the energies E and E 0 ; the integral over d3 p integrates over the electron distribution, and the
factor
d
d
(2.88)
d
specifies the probability for scattering photons from frequency
to frequency 0 or backward;
we assume thermal photon and electron distributions, and restrict
ourselves to the limit of Thomson scattering, which applies if
h  mc2 ;

(2.89)

moreover, we assume small changes in the photon frequency,


hence
0  ;
(2.90)

CHAPTER 2. SCATTERING

25

moreover, the electron energy distribution is


!
E
N(E) exp
,
kT e

(2.91)

and energy conservation requires


E 0 = h h ;

(2.92)

now, both n() and N(E) can be expanded in Taylor series up to


second order,
n
1 2 n 2
+ O(3 ) ,
+
(2.93)

2 2
N
1 2 N 2 2
0
h + O(3 ) ,
N(E ) = N(E)
h +
2
E
2 E
where (2.91) allows us to use
n(0 ) = n() +

N
N(E)
=
,
E
kT e

N(E)
2 N
=
;
2
E
(kT e )2

(2.94)

for simplification, we now define the dimension-less photon energy, scaled by the thermal electron energy
x

h
kT e

(2.95)

and find
1 2 n 2
n
x ,
n(x0 ) n(x) + x +
2
x
2
x
#
"
x2
0
N(E ) N(E) 1 + x +
;
2

(2.96)

with these approximations, we return to the original equation


(2.86) for n() and obtain
"
#
n
n
=
+ n(n + 1) I1
t
x
"
#
n
1 2 n
+ 2(1 + n) + n(n + 1) I2 ,
+
(2.97)
2 x2
x
with the abbreviations
Z
Z
d
3
Ii
d p d
cxi N(E)
d

(2.98)

the energy change of a photon scattering off a moving electron


follows from (2.71), adopting the non-relativistic limit
p2
E = mc +
2m
and using (2.89) and (2.90); this yields
2

h =

h
(~e ~e0 ) ~p
mc

x =

(2.99)

x
(~e ~e0 ) ~p ; (2.100)
mc

CHAPTER 2. SCATTERING

26

using this result, the integrals Ii can be carried out straightforwardly; with the unpolarised Thomson cross section (2.34), we
first find
kT e x2
;
(2.101)
I2 = 2T ne c
mc2
for evaluating I1 , we note that I1 is the mean rate of relative energy transfer, quantified by x from the electrons to the photons,
and therefore the mean energy transfer rate, divided by kT e ; from
(2.77), we know that this is
D

E x(kT e )2
(4 x)
E =
me c2

(2.102)

per scattering, and multiplying with the collision rate ne T c gives


I1 =

kT e
ne T c x(4 x) ;
me c2

(2.103)

with these two expression for Ii , we find the time derivative of n


to be
"
!#
me c2 1 n
1 4 n
2
x
= 2
+n+n
;
(2.104)
kT e ne T c t
x x
x
we finally transform the time t to the Compton parameter, using
dy =

kT e
ne T c dt
me c2

to find the Kompaneets equation


"
!#
n
1 4 n
2
x
=
+n+n
;
y x2 x
x

(2.105)

(2.106)

the hot gas in galaxy clusters is much hotter than the cosmic background radiation; then, we can approximate the right-hand side of
(2.106) to lowest order in x,
2 n
n
n
x2 2 + 4x ;
y
x
x

(2.107)

inserting here the occupation number in thermal equilibrium, n


(e x 1)1 , we find
!
n
x2 e x (1 + e x )
4xe x
(2.108)
= y
x
n
(e x 1)2
e 1
for the relative change of the occupation number, where x is now
h/kT and no longer h/kT e !

Chapter 3
Radiation Transport and
Bremsstrahlung
3.1

Radiation Transport Equations

we start with the collision-less Boltzmann equation for describing


the temporal change of the photon distribution function in phase
space,
n ~ n n
+ ~p
=0,
(3.1)
+
t
~x
~p
which is valid in absence of collisions;
for photons, we have ~v = c~e, where ~e is the unit vector in the
direction of light propagation; moreover, ~p = 0 in absence of
systematic external forces (such as gravitational lensing); since
the intensity I is proportional to n, the Boltzmann equation for
photons can also be written as
1 I
I
+ ~e
=0;
c t
~x
we now define the following quantities:
Z
Z
1
F~
d ~e I , P,i j
d ei e j I
c
and recall the spectral energy density
Z
1
U =
d I ;
c

(3.2)

(3.3)

(3.4)

integrating the Boltzmann equation (3.2) first over d, we obtain


the equation
U ~ ~
+ F = 0 ;
(3.5)
t
27

further reading: Shu, The


Physics of Astrophysics, Vol I:
Radiation, chapters 2, 3, and
15; Rybicki, Lightman, Radiative Processes in Astrophysics,
chapter 5; Padmanabhan, Theoretical Astrophysics, Vol. I: Astrophysical Processes, sections
6.86.9

CHAPTER 3. RADIATION TRANSPORT AND BREMSSTRAHLUNG28


which has the form of a continuity equation and identifies F~ as
the spectral radiation current density (spectral because it retains
the dependence on frequency ); this equation expresses energy
conservation in the radiation field;
if we multiply (3.2) with ei first before integrating over d, we
find
Z
Z
I
1
I
d ei
=0,
(3.6)
+ d ei e j
c
t
x j
and hence

P,i j
1 F,i
+c
=0;
c t
x j

(3.7)

this equation describes the change of the momentum current density, because
U 

(c~e)
(3.8)
c
is the momentum density of the radiation field, and thus
Z
1 F~
1
d I~e
(3.9)
= 2
c t
c t
is c times the temporal change of the momentum current density;
Eq. (3.7) expresses momentum conservation;
in presence of emission, stimulated emission and absorption, we
know from the first chapter that the energy equation must be augmented by source and sink terms on its right-hand side; we had
dI
1 dI
= j I =
;
dl
c dt

(3.10)

integrating over d, and assuming that j and are isotropic, we


find
dU
= 4 j U c = 4 j cU ;
(3.11)
dt
we now re-define the emissivity,
4 j j0 j ,

(3.12)

i.e. we refer it to the mass density, and write


dU
= ( j cU ) ;
dt

(3.13)

likewise, the momentum-conservation equation


1 dI
= j I
c dt

(3.14)

becomes after multiplication with ~e and integration over d


Z
Z
Z
1d
d ~e I =
d j~e d I~e ,
(3.15)
c dt

CHAPTER 3. RADIATION TRANSPORT AND BREMSSTRAHLUNG29


and thus

1 dF~
= F~ = F~ ,
(3.16)
c dt
where we have assumed again that j and are isotropic
including the emission and absorption terms, the transport equations are modified to read
U ~ ~
+ F = ( j cU )
t
P,i j
1 F,i
+c
= F,i ;
c t
x j

(3.17)

these equations do not contain scattering terms yet!


since the change in the momentum current density corresponds to
a force density, and this force is caused by the interaction between
radiation and matter, an oppositely directed and equally strong
force must act on the matter as radiation pressure force; thus
Z
~
~
(3.18)
frad =
F d
c 0
is the density of the radiation pressure force;

3.2

Local Thermodynamical Equilibrium

the moment equations for U and F~ are by no means easier to


handle than the Boltzmann equation whose moments they are;
we obviously need an additional approximation, or condition, in
order to close the moment equations; the closure means that
they can then be solved without progressing indefinitely to higher
orders of moments;
often, the mean free path of the photons is much smaller than
the dimensions of the system under consideration; then, we can
assume that thermodynamical equilibrium is locally established
between the radiation field and the matter; under this condition,
I B (T ) ,

(3.19)

i.e. the specific intensity of the radiation field is the Planckian


intensity of a black body, and
Z
1
4
U =
d I =
B (T ) ;
(3.20)
c
c

CHAPTER 3. RADIATION TRANSPORT AND BREMSSTRAHLUNG30


under such circumstances, there is obviously no radiation flux any
more because the radiation field is isotropic; in order to estimate
the flux nonetheless, we study the orders of magnitude of the different terms in the moment equations;
time derivatives can typically be neglected because temporal
changes of the quantities U , F~ and P,i j occur on an evolutionary
time scale, while the other terms change according to the streaming of the photons, thus approximately on time scales of order
(mean free path)/c;
if we first ignore F~ /t, we obtain
F,i

c P,i j
x j

(3.21)

in the approximation of Local Thermodynamical Equilibrium


(LTE), we further have
P,i j P i j =

U
i j
3

(3.22)

because of the (local) isotropy of the radiation field, and thus


F,i

c U
c U

,
xi
R

(3.23)

where R is a typical dimension of the system; the mean free path


is determined by
n = 1 ,
and (3.23) can thus be approximated by
 

,
|F,i | cU
R

(3.24)

(3.25)

which is smaller by a factor /R compared to the transparent


case (in which 0 and R;
using this estimate for F~ , we return to the Eq. (3.17) for the partial time derivative of U ; as before, we ignore the time derivative,
such that the only term remaining on the left-hand side is
~ F~ cU ;

R R

(3.26)

the second term on the right-hand side is


c
R
cU U

!2

~ F~ 
~ F~ ;

(3.27)

CHAPTER 3. RADIATION TRANSPORT AND BREMSSTRAHLUNG31


thus, because of the assumption of local thermodynamical equilibrium, the divergence of F~ is negligibly small; consequently,
we must require
j cU

4
=
B (T ) ,
c
c

(3.28)

as anticipated;
accordingly, if  R and tevol  /c, the solutions of the
moment equations are
F,i

c P,i j
,
x j

4
B (T ) ;
c

(3.29)

because of (local) isotropy, we had


P,i j
and thus

4
U
i j
B (T ) i j ,
3
3c

!
4 B T
F,i
,
3 T xi

(3.30)

(3.31)

i.e. the flux will become proportional to the temperature, which is


characteristic for diffusion processes;
for convenience, we now introduce the Rosseland mean opacity,
R  B (T ) 

d 1 T
0
(3.32)
R1 R  B (T )  ;

d
T
0
here, we can use the fact that
!
!
Z
Z
B (T )

caT 4
d
=
,
d B (T ) =
T
T 0
T 4
0

(3.33)

where

2 k4
erg
= 7.6 1015 4 3
3
15 (~c)
K cm
is the so-called Stefan-Boltzmann constant;
a

using this, we obtain the expression


Z
Z
4
T
d B
F~ =
dF~ =
3 ~x 0 T
0
!
4 ~ d caT 4
c ~  4
=
=
T
aT
3R
dT 4
3R
for the radiative energy flux;

(3.34)

(3.35)

CHAPTER 3. RADIATION TRANSPORT AND BREMSSTRAHLUNG32


the energy which is streaming away interacts with the absorbing
matter and thus exerts a force on it, which is determined by the
right-hand side of the momentum-conservation equation, as described above:
Z
Z
4

B T

d F~ =
d
f~rad =
c 0
c 3 0
T ~x
1~
4
~ ,
= (aT
(3.36)
) = P
3
which equals just the negative pressure gradient;
a remark on units: the unit of U is
[U ] =

erg
,
cm3 Hz

(3.37)

cm2
,
g

(3.38)

the unit of is
[ ] =
and thus the unit of F~ is

[F~ ] = [c][U ] =

erg
,
cm2 t Hz

(3.39)

and the unit of f~rad is


g s cm2
erg
erg
(3.40)
Hz
=
cm3 cm g cm2 s Hz
cm4
g cm2 1
g cm 1
dyn
force
=
= 2
=
=
,
4
4
3
3
s cm
s cm
cm
volume

[ f~rad ] =

as it should be;

3.3

Scattering

so far, we have only considered emission and absorption, but neglected scattering; scattering changes the distribution function of
the photons by exchanging photons with different momenta; if we
assume for simplicity that the scattering process changes the photons momentum, but not its energy, we can write the scattering
cross section in the form
d(~e ~e0 )
= (~e, ~e0 ) ,
d

(3.41)

where ~e and ~e0 are unit vectors in the propagation directions of


the incoming and the outgoing photon; the function (~e, ~e0 ) is
normalised, symmetric in its arguments and dimension-less and
describes the directional distribution of the scattered photons;

CHAPTER 3. RADIATION TRANSPORT AND BREMSSTRAHLUNG33


scattering increases the distribution function n(~e) according to
#
"
Z


dn(~e)
=
d0
Ne c(~e, ~e0 )
n(~e0 ) [1 + n(~e)] ,
|
{z
}
dt +
# of scatterings ~e ~e0
(3.42)
where the factor [1 + n(~e)] is included for describing stimulated
emission of photons with momentum direction ~e;
analogously, losses due to scattering are given by
#
"
Z


dn(~e)
=
d0 Ne c(~e, ~e0 ) n(~e) [1 + n(~e0 )] ,
dt

(3.43)

and thus the total change of n(~e) due to scatterings becomes


Z


dn(~e)
=
d0 Ne c(~e, ~e0 )
dt


n(~e0 ) [1 + n(~e)] n(~e) [1 + n(~e0 )]
Z



=
d0 Ne c(~e, ~e0 ) n(~e0 ) n(~e) , (3.44)
in which the terms from stimulated emission cancel exactly;
since the integral over the solid angle only concerns the direction
of ~e0 , we obtain from (3.44)
Z
n(~e)
= Ne cn(~e) + Ne c d0 (~e, ~e0 )n(~e0 ) ,
(3.45)
dt
and thus
Z
1 dn(~e)
sca
sca
= n(~e) +
d0 (~e, ~e0 )n(~e0 ) ,
(3.46)
c dt
where we have introduced the scattering opacity through sca =
Ne ;
since the intensity at fixed frequency is proportional to the occupation number, the same equation (3.46) also holds for I ; therefore, the transport equation for the specific intensity is changed in
presence of scattering to
1 I
I
1 dI
=
+ ~e
(3.47)
c dt
c t
~x
"
#
Z
j
abs
sca
0
0
0
=
I I d (~e, ~e )I (~e ) ;
4
again, we now take the moments of the transport equation in order
to see how the moment equations are changed by scattering; the
first moment is obtained by integrating (3.48) over d,
U ~ ~
+ F = j abs cU sca cU + sca cU (3.48)
t
due to the normalisation of (~e, ~e0 ); therefore, the scattering terms
cancel, and the equation for U remains unchanged;

CHAPTER 3. RADIATION TRANSPORT AND BREMSSTRAHLUNG34


the next moment equation simplifies if we further assume that
Z
Z
0
d(~e, ~e )~e = 0 =
d0 (~e, ~e0 )~e0 ,
(3.49)
which holds for many scattering processes (e.g. Thomson scattering); then, the second moment equation reads
P,i j
1 F,i
+c
= abs F,i sca F,i
c t
x j
= (abs + sca )F,i ,

(3.50)

i.e. the scattering opacity is simply added to the absorption opacity here; with a suitable modification of the Rosseland mean opacity, the diffusion approximation remains valid which we have obtained above;

3.4
3.4.1

Bremsstrahlung
Spectrum of a Moving Charge

a radiation process which is very important in astrophysics is due


to electrons which are scattered off ions and radiate due to the
acceleration they experience; in order to describe it, we start again
from Larmors equation, which says
dE 2e2
= 3
dt
3c

2
~x ,

(3.51)

where e is the charge of the accelerated particle (the electron, in


most cases), ~x is its acceleration, and dE/dt is the power radiated
away;
as a function of frequency, this equation can be written as follows:
Z
2e2 2
2e2 2
dE = 3 ~x
E= 3
dt ~x ;
(3.52)
3c
3c
if we Fourier-transform the particles trajectory,
Z
Z
d
it
~x() =
~x()eit ,
dt ~x(t)e , ~x(t) =

2
we first have
Z
d h 2 i it
~x =
~x() e
2

~x = 2 ~x() ,

and we can use Parsevals equation,


Z
Z
d 2
2
f () ;
dt f (t) =

(3.53)

(3.54)

(3.55)

CHAPTER 3. RADIATION TRANSPORT AND BREMSSTRAHLUNG35


combining these results yields
Z
Z
2e2 d 2 4e2 d 4 2
E= 3
~x() ,
~x() = 3
3c 2
3c 0 2

(3.56)

and, by differentiation,
dE
2e2 4 2
~x() ;
=
d 3c3

(3.57)

this is a general expression valid for all radiation processes; in


order to make progress, we need the Fourier transform of the specific particle trajectory;

3.4.2

Hyperbolic Orbits

classically, the electron follows a hyperbola around the ion in the


orbital plane perpendicular to the (conserved) angular momentum; the focal point of the hyperbola is the centre of mass, which
we assume to coincide with the centre of the scattering ion, i.e. we
neglect the mass of the electron; in polar coordinates, the trajectory is given by
p
,
(3.58)
r() =
1 +  cos
with the parameters
Lz2
p=
= a( 2 1)
m

2Lz2 E
and  1 + 2
m

!1/2
,

(3.59)

where Lz = bmv is the angular momentum in z direction, v is


the initial velocity at infinity, E = mv2 /2 is the kinetic energy at
infinity and thus the total energy, and quantifies the coupling
strength; for electrons orbiting nuclei at rest with charge Ze,
= Ze2

(3.60)

as for solving the Kepler problem, we introduce a parameter


(the eccentric anomaly), of which we require that it satisfy
r = a( cosh 1) ;

(3.61)

we find the relation between and by inserting (3.61) into


(3.58), using (3.59)
a( 2 1)
= a( cosh 1)
1 +  cos
if we want = 0 where = 0;

cos =

 cosh
,
 cosh 1
(3.62)

CHAPTER 3. RADIATION TRANSPORT AND BREMSSTRAHLUNG36


energy conservation implies that the time t when the electron
reaches the distance r from the scattering ion is
Z r
Z r
xdx
dx
=
t=


1/2 ,
 

1/2
Lz2
Lz2
r0
r0
2

2
2
E + x 2mx2
Ex + x 2m
m
m
(3.63)
where
Lz2 = ma( 2 1)
(3.64)
was used from (3.59); furthermore, we have
a=

Lz2 2 m
p

=
=
2
2
 1 m 2Lz E 2E

E=

;
2a

combining, we first find


r Z r
m
xdx
t=
,
h
2 r0 x2 + x a ( 2 1)i1/2
2a
2

(3.65)

(3.66)

which can be transformed with (3.61) to obtain


r
r
Z
ma3
ma3
( cosh 1)d =
t=
( sinh ) ; (3.67)
0

the coordinates x and y in the orbital plane satisfy


 cosh
x = r cos = a( cosh 1)
= a( cosh )
 cosh 1

(3.68)
y = r sin = a  2 1 sinh ,
where we have used (3.61) and (3.62); with these expressions, we
return to the Fourier transform of x and y
since
x = i x() ,

(3.69)

we have
Z
i
i
x() =
x() =
dt x(t)eit


Z
dx d it()
i
dt
e
=
d dt
Z
ia
=
d sinh eit() ;

using (3.67), we can write
r

ma
it()
( sinh ) ;
e
= exp i

(3.70)

(3.71)

CHAPTER 3. RADIATION TRANSPORT AND BREMSSTRAHLUNG37


from (3.65), we find
a=

,
mv2

(3.72)

m2

= 3 ,
3
6
m v mv

(3.73)

moreover,
r

and thus

ma3
=

eit() = ei( sinh )

(3.74)

putting this result into (3.70), we write


Z
ia
x() =
d sinh ei( sinh )

Z
ia  2 1
y () =
d cosh ei( sinh ) ; (3.75)

these expressions can be analytically integrated and lead to firstorder Hankel functions, which will not be discussed in detail here;
forming
|~x()|2 = x() x () + y ()y ()

(3.76)

and inserting the result into (3.57) yields the desired


bremsstrahlung spectrum;

3.4.3

Integration over the Electron Distribution

having obtained the spectrum dE/d for a single charge, we now


have to integrate over a distribution of charges; we do this by integrating over all impact parameters b from 0 to after multiplying
dE/d with
ni ne v 2bdb ,
(3.77)
which is the number of scatterings per unit volume and unit time
between ions and electrons with the number densities ni and ne ,
respectively; for a fully ionised pure hydrogen gas, ni = ne n;
preparing the integration, we note from (3.59) and (3.72) that
2 = 1 +

b2 m2 v4
b2
=
1
+
,
2
a2

(3.78)

such that the integration over b can be transformed into an integration over ,
d = 

bdb bdb
= 2
a2
a

where  [1, ) while b [0, );

bdb = a2 d ,

(3.79)

CHAPTER 3. RADIATION TRANSPORT AND BREMSSTRAHLUNG38


inserting suitable approximations for the first-order Hankel functions, we find after carrying out the b integration and inserting
(3.60)
 


2 mv3
mv3

dE
16Z 2 e6 n2
ln Ze2   Ze2 
,
(3.80)


dVdtd
3m2 c3 v
 mv
Ze2
3
where 1.78 is Eulers constant;
we write this result as
16Z 2 n2 e6 1
dE
=
gff (v, ) ,
dVdtd
3 3m2 c3 v

(3.81)

introducing the so-called gaunt factor gff , which usually depends


at most weakly on v;
in a dilute thermal plasma, the electrons have a Maxwellian velocity distribution, but for emitting a photon of energy ~, an
electron needs at least an energy of
r
m 2
2~
vmin = ~ vmin =
;
(3.82)
2
m
the thermal average of the inverse velocity is then
* +
 m 3/2 Z
1
1
2
4v2 dv emv /2kT
=
v
2kT
v
vmin
r
2m ~/kT
=
e
;
kT

(3.83)

replacing 1/v in (3.81) by the average (3.83) finally yields the


specific emissivity of a thermal plasma due to bremsstrahlung,
r
dE
2m ~/kT
16Z 2 n2 e6
j() =
e
gff (v, ) ; (3.84)
2
3
dVdtd
kT
3 3m c
the volume emssivity is the integral of j() over frequency ,
r

dE
16Z 2 n2 e6
2mkT
2
T;
(3.85)
j=
=
g
(v,
)

n
ff
dVdt
3~m2 c3

numerically, the volume emissivity in cgs units is

!1/2

kT

20

6.69 10

erg

 n 2
!1/2

2
kT
24
j = Z gff

2.68 10

cm3

keV

 T 1/2

28

7.86 10
K

(3.86)

CHAPTER 3. RADIATION TRANSPORT AND BREMSSTRAHLUNG39


as an example, we consider the X-ray emission of a massive
galaxy cluster with kT = 10 keV; typical clusters reach electron
densities of n 103 cm3 in their cores; let us assume for simplicity that the X-ray emitting gas with that electron density fills
a volume of 1 Mpc3 ;
assuming fully ionised pure hydrogen, we put Z = 1 and gff = 1
for simplicity; then, (3.86) yields

LX V j (3.1 1024 )3 106 2.68 1024 10


2.5 1044 erg s1 ,
(3.87)
which makes galaxy clusters the most luminous X-ray sources on
the sky;
with an average energy of 10 keV = 1.6e 8 erg per photon,
this luminosity corresponds to
NX 1.6 1052 s1

(3.88)

photons emitted by the cluster per second; if the cluster is at a


distance of, say, 100 Mpc 3.1 1026 cm, these photons are distributed over an area of 1.2 1054 cm2 , such that an X-ray detector with a typical collecting area of a few hundred cm2 sees
100

1.6 1052 1
s 1 s1
1.2 1054

(3.89)

i.e. this enormous X-ray luminosity produces a flux of approximately one photon per second in a typical X-ray detector;

Chapter 4
Synchrotron Radiation,
Ionisation and Recombination
4.1
4.1.1

Synchrotron Radiation
Electron Gyrating in a Magnetic Field

a further very important radiation process is the emission of radi~ in such a field,
ation by electrons moving in a magnetic field B;
electrons spiral around field lines, with their angular frequency
given by
ceB
eB
B =
=
,
(4.1)
E
mc
where E is the electron energy, and is the usual Lorentz factor;
numerically, we have
 B 
1
,
(4.2)
B 17.6 MHz
1G
i.e. synchrotron radiation is typically emitted at radio frequencies;
the radius of the projection of the spiral orbit perpendicular to the
magnetic field is
v
mcv
rB =
=
,
(4.3)
B
eB
~ superand the complete motion is the circular motion around B,
~
posed by a drift along B;
we employ Larmors equation
dE 2e2
= 3
dt
3c

40

2
~x

(4.4)

further reading: Shu, The


Physics of Astrophysics, Vol I:
Radiation, chapters 18, 19, 21
23; Rybicki, Lightman, Radiative Processes in Astrophysics,
chapters 6 and 10; Padmanabhan, Theoretical Astrophysics,
Vol. I: Astrophysical Processes,
sections 6.106.12

CHAPTER 4. SYNCHROTRON RADIATION, IONISATION AND RECOMBINATION41


again to calculate the emission; assuming E~ = 0, the Lorentz
force is
e
e
~
~ ;
F~L = (~v B)
~x =
(~v B)
(4.5)
c
mc
~ then |~v B|
~ = vB sin
let now be the angle between ~v and B,
and
 e 2
2

|~x| =
(vB sin )2 ,
(4.6)
mc
and, from Larmors equation (4.4),
2e4 2 v2 B2 sin2
dE
=
(4.7)
dt
3m!2 c5
2 e2
B2 2 2 2
2
2 2 2
c

B
sin

=
2c
=
sin ,
T
3 mc2
8
where we have identified the Thomson scattering cross section;
averaging over all pitch angles , finding
Z
1 3
2
2
hsin i =
sin d = ,
2 0
3

(4.8)

we obtain the radiation energy emitted per unit time by an


isotropic electron distribution in a magnetic field,
dE 4 B2
4
=
cT 2 2 = UB cT 2 2 ,
dt
3 8
3

(4.9)

where we have inserted the energy density UB = B2 /(8) of the


magnetic field;

4.1.2

Beaming and Retardation

we had seen in Sect. 2 of Chap. 1 that solid angles are deformed


by Lorentz transformations according to
d0 =

d
;
2 (1 cos )2

(4.10)

relativistically moving charges thus focus their emission into their


forward direction; for estimating the opening angle of the resulting cone, we require that
1
1

,
1 cos 2(1 )

(4.11)

i.e. we want to estimate the half-width of the cone; approximating


the cosine in (4.11) to second order,
1+

2
2(1 ) ,
2

(4.12)

CHAPTER 4. SYNCHROTRON RADIATION, IONISATION AND RECOMBINATION42


thus
2 2(1 ) (1 + )(1 ) = 2

(4.13)

where we have approximated 2 (1 + ) for relativistic particles,


and
1

(4.14)

due to the narrow emission cone with the opening angle 1 ,


an observer sees an electron spiralling in a magnetic field only
during the short moment while the cone is moving past him;
moreover, this means that the radiated electric field can depend
on only through the combination , because a change in
causes an immediate change in ;
finally, the arrival time of the radiation at the observer depends
also on the angle
we now consider two radiation signals which leave the electron at
times t1 and t2 = t1 + t; if the velocity of the electron is v and its
orbital radius is rB , we have
rB = vt

t=

rB
;
v

(4.15)

the signal leaving at t2 needs less time to get to the observer,


namely by the amount
rB
;
(4.16)
c
this is of course the usual retardation due to the finite light speed;
the observing time for the second signal is thus
rB rB 
v
tobs = t
1
;
(4.17)
=
c
v
c
now we obtain for the angle, where the electron was when it emitted the radiation which arrives at the observer at time tobs
vtobs 
v 1
obs =
1
;
(4.18)
rB
c
approximating the squared Lorentz factor by
v2
= 1 2
c
2

!1

v 1 
v 1 1 
v 1
= 1
1+

1
c
c
2
c


(4.19)

for v c, we finally find


obs =

2vtobs 2
,
rB

(4.20)

CHAPTER 4. SYNCHROTRON RADIATION, IONISATION AND RECOMBINATION43


from which we obtain
obs


 4
2vtobs 3
= tobs 2B 3 sin c tobs ,
rB
3

(4.21)

because v = rB B sin ; moreover, the last equation defines the


angular frequency

3
3  eB
3
c
B sin =
sin 2
2
2
mc
!
B  E 2
,
(4.22)
100 MHz
G GeV
where the factor 3/2 was introduced for later convenience;

4.1.3

Synchrotron Spectrum

since the electric field can depend on only through the factor
and depends only through the factor c tobs on the observing time, a Fourier transform must find that the spectrum of
synchrotron radiation can depend on frequency only through
the ratio

;
(4.23)
c
qualitatively, synchrotron radiation is determined by the following properties:
the basic frequency for 1 is the cyclotron frequency
cyc =

eB
;
mc

(4.24)

for higher velocities, overtones of the cyclotron frequency


will be added whose amplitudes will decrease by powers of
v/c;
since B 1 , the electrons angular orbital frequency will
decrease as v/c increases;
finally, the radiation is limited to a cone which allows the
observer to see the radiation only during time intervals
t 3 1
B ;

(4.25)

therefore, we expect a spectrum which consists of a sequence of


sharp maxima at B and its overtones and cuts off at
c

1
B 3 ;
t

(4.26)

CHAPTER 4. SYNCHROTRON RADIATION, IONISATION AND RECOMBINATION44


while is increasing, the contribution due to the overtones will
increase, and the maxima will be broadened due to the electron
distribution in the Lorentz factor and the pitch angle ; thus, a
continuum is formed, which ends near c ;
most of the energy will be emitted near c ; since
c B 3 B2 BE 2
the electron energy E is proportional to
r
c
E
,
B

(4.27)

(4.28)

more precisely
E 4c1/2 B1/2 erg ,

(4.29)

where c = c /(2) is and all quantities are expressed in cgs units;


synchrotron radiation is strongly linearly polarised because is results from the orbital motion of electrons perpendicular to the
magnetic field; a circular polarisation parallel to the magnetic
field is effectively averaged out by the pitch-angle distribution;

4.2
4.2.1

Photo-Ionisation
Transition Amplitude

many astrophysical radiation processes are accompanied by ionisation and recombination; in ionising processes, a sufficiently energetic photon removes an electron from a bound state and places
it into an unbound state, and the reverse process happens in recombinations;
as an example, we consider the cross section for ionising a hydrogen atom from its ground state; for doing so, we first need
to compute the transition rate between bound and unbound states
which is caused by a perturbing photon;
according to the quantum-mechanical perturbation theory, the
transition probability within a time interval t is given by
P(t) = |a2 (t)| ,
where a is the probability amplitude
Z t
dt 0
hE |H|Ei ,
a(t) =
~
0

(4.30)

(4.31)

CHAPTER 4. SYNCHROTRON RADIATION, IONISATION AND RECOMBINATION45


and |Ei and |E 0 i are the initial and final states of the electron,
respectively, which are the bound and unbound states here; H is
the Hamiltonian of the perturbing interaction; the states |Ei and
|E 0 i are written as eigenstates of the Hamiltonian;
the canonical momentum of a particle in an electromagnetic field
can be written as
~,
~ = ~p e A
(4.32)
P
c
~ is the vector potential; accordingly, the total Hamiltonian
where A
is
~2
P
1 

~p
H=
=
2m 2m

 e 2 A
~2
~p2
e ~ 2
e ~
A =
A ~p +

; (4.33)
c
2m mc
c 2m

the Hamiltonian of the interaction between the electron and the


field is thus
e ~
H =
A ~p ;
(4.34)
mc
~ satisfies the transversal (or Coulomb)
we have assume here that A
~ A
~ = 0; then, A
~
gauge, in which the scalar potential = 0 and
satisfies the wave equation
!
1 2 ~ 2 ~
A=0;
(4.35)
c2 t2
~ x, t) is decomposed into plane waves,
if A(~
Z
d3 k ~
~
~
A(~x, t) =
Ak (t)eik~x ,
3
(2)

(4.36)

the wave equation implies for the Fourier amplitudes


~ k (t) + c2 k2 A
~k (t) = 0
A

(4.37)

~k (t) are independent of each


because the individual amplitudes A
~k (t) satisfies an oscillator equation and can be
other; therefore, A
written in the form
~k (t) = A~e eit ;
A
(4.38)
the dispersion relation = kc holds here, ~e is the unit vector in
polarisation direction, and Ak is the amplitude of the given mode
of the vector potential;
the interaction Hamiltonian between field and electron is thus
H =

eA i(t~k~x)
~e ~p ;
e
mc

(4.39)

CHAPTER 4. SYNCHROTRON RADIATION, IONISATION AND RECOMBINATION46


the transition matrix element hE 0 |H|Ei can now be calculated as
follows: first, the state are represented by the wave functions
|Ei = (~x)eiEt/~ ,

|E 0 i = 0 (~x)eiE t/~ ,

(4.40)

in which (~x) and 0 (~x) are the spatial amplitudes, and the phase
factors describe the time evolution; using these, the probability
amplitude (4.31) turns into

Z t Z
~
0
A
e eik~x

3
0
~e ~p (~x) ei(E /~E/~)t ;
a(t) =
dt d x (~x)
~c 0
m
(4.41)

4.2.2

Transition Probability

we abbreviate the notation by introducing the matrix element between the initial and the final state,
~ Z
e eik~x
e
~
Mfi
d3 x 0 (~x) ~p (~x) ,
(4.42)
m
m
in terms of which we can write the probability amplitude (4.41)
as
Z t
A
0
~
(~e Mfi )
a(t) =
dt ei(E /~E/~)t ;
(4.43)
~c
0
the time integral can be evaluated as follows:
Z t
Z t
E0 E
it
dt e =
dt (cos t+i sin t) ,  ; (4.44)
~ ~
0
0
taking the absolute square of this expression, which we shall later
need, we find
Z t

!2

sin(t/2)
it
;
(4.45)
dt e =


/2
0
in the limit of t this expression approaches a function,
sin2 (t/2)
lim
2tD () ;
t
(/2)2

(4.46)

combining these results, the transition probability (4.31) turns out


to be

2
2A2 t

~
P(t) =
D () ~e Mfi ,
(4.47)
~2 c2
and thus the transition rate is
!
~ 2
P(t) 2A2
E0 E
R=
= 2 2 D
~e M
(4.48)
fi ;
t
~c
~
~

CHAPTER 4. SYNCHROTRON RADIATION, IONISATION AND RECOMBINATION47


we further introduce the abbreviation
E0 E
fi
,
~

(4.49)

i.e. we assign an angular frequency to the energy difference E 0 E


between the final and initial states;
what remains is the determination of the field amplitude Ak and
~ fi ; we first consider the amplitude Ak ;
the matrix element M
the electric field belonging to the vector potential
~k (t) = Ak ~e ei(t~k~x)
A

(4.50)

is

~k (t)
1 A
iAk~e i(t~k~x)
E~ k =
=
e
,
(4.51)
c t
c
and the energy density in an electromagnetic field in vacuum can
be written as
~ 2 E~ 2
E~ 2 + B
U=
=
;
(4.52)
8
4
similarly, the energy density in photons of angular frequency is
1 N~
,
2 V

(4.53)

where the factor 1/2 appears because the two independent polarisation directions need to be distinguished; we set N = 1 so that
we shall only have to multiply with the number of photons in the
volume V later; thus, we find by comparing (4.51), (4.52) and
(4.53)
r
2~c2
N~ A2k 2
=

A
=
(4.54)
k
2V
4c2
V
for a single photon (N = 1); the unit of Ak is
erg s cm2 s
[Ak ] =
cm3 s2

!1/2
=

 erg 1/2
cm

g1/2 cm1/2
,
s

from which we find for the unit of the electric field



g1/2
~
= [Ak ] cm1 =
,
[Ek ] = [Ak ]
c
cm1/2 s

(4.55)

(4.56)

as it must be;
for the transition rate (4.48), we now have
42 ~ 2
R=
~e Mfi D (fi ) ;
~V

(4.57)

CHAPTER 4. SYNCHROTRON RADIATION, IONISATION AND RECOMBINATION48


according to our choice of the amplitude Ak , this is the transition
rate caused by a single photon; we must now multiply the expression (4.57) with the number of available photons, which is
Vd3 k
n()
,
(2)3

(4.58)

in which n() is the occupation number of photons with frequency and polarisation state ~e, i.e. the photon number density
per phase-space cell;
since = ck, the volume element in k-space is
d3 k = k2 dkd =

2 d
d ,
c3

(4.59)

and we obtain the number of ionisation transitions per unit time,


unit frequency and unit solid angle as
dP
n() ~ 2
=
(4.60)
~e Mfi D (fi ) ;
dtdd 2~c3

4.2.3

Transition Matrix Element

~ fi as defined in
finally, we have to calculate the matrix element M
(4.42); if the wave length of the incoming light, = 2/k, is much
larger than the extent of the wave function of the bound electron,
we can approximate
~
eik~x 1
(4.61)
and are left with
~ fi
M

d3 x 0 (~x)

e
~p (~x) ;
m

(4.62)

this approximation is called dipole approximation for the following reason: the momentum operator ~p can be expressed by
the commutator
~p =

im
im
[H, ~x] = (H ~x ~xH) ;
~
~

(4.63)

~ fi (4.62) can be transformed to


therefore, the matrix element M
Z
e
im
0
~ fi
M
(E E)
d3 x 0 (~x) ~x (~x) = ifi e~x ,
(4.64)
m ~
i.e. it turns into a dipole matrix element;
during the transition from the bound to the free state, the Hamiltonian changes, and thus it is preferable in this context to use the
momentum operator instead; we now insert the wave function of

CHAPTER 4. SYNCHROTRON RADIATION, IONISATION AND RECOMBINATION49


the ground state in the hydrogen atom as the wave function for
the initial state,
er/a0
(~x) = q
,
(4.65)
3
a0
where a0 is Bohrs radius
a0

~2
= 4.7 108 cm ,
me2

(4.66)

while the free electron is described by the plane wave


~

eike ~x
(~x) = ,
V
0

(4.67)

where ~ke is the wave vector of the free electron, which is related
to the momentum by ~pe = ~~ke ;
we first confirm that the dipole approximation can be applied
here; for short-wave light, 1000 = 105 cm, which is
almost three orders of magnitude larger than Bohrs radius a0 ,
 a0 ;
now, the transition probability between the initial and the final
state equals the reverse transition rate,

2
final|~e ~p|initial = initial|~e ~p|final ,
(4.68)
~ we find
and inserting the momentum operator ~p = i~,


~~e ~ke
initial|~e ~p|final = q
Va30

d3 x ex/a0 eike ~x ;

(4.69)

denoting the angle between ~ke and ~x by , we can write


Z
Z
Z
3
x/a0 i~ke ~x
2
d xe
e
= 2
x dx
sin d ex/a0 eike x cos
0
0
Z
Z 1
= 2
x2 dx
d(cos ) ex/a0 eike x cos
1
Z0
sin ke x
= 4
x2 dxex/a0
ke x
0
3
8a0
,
(4.70)
=
(1 + ke2 a20 )2
which turns the square of the transition matrix element into

2 64~2 e2 a3 |~e ~k |2
e
0
~e M
~ fi =
,
2
Vm
(1 + ~ke2 a20 )4

(4.71)

CHAPTER 4. SYNCHROTRON RADIATION, IONISATION AND RECOMBINATION50


and the number of transitions per unit time, unit frequency and
unit solid angle (4.60) becomes
dP
n() ~ 2
=
(4.72)
~e Mfi D (fi )
dtdd
2~c3
32~e2 n()a30 |~e ~ke |2
=
D (fi ) ;
Vm2 c3
(1 + ~ke2 a2 )4
0

4.2.4

Cross Section

the cross section for photo-ionisation is determined by the relation


dP
dP Vd3 pe dP Vd3 ke
=
dNe =
dt
dt (2~)3
dt (2)3
d3 p
!
;
= () cn()
(2~)3

(4.73)

this means that the flux of incoming photons, cn() times the
number density of states in phase space, multiplied with the cross
section, is the number of transitions per unit time; this must equal
the transition rate times the number of available final states for
the electron;
in order to obtain the cross section, we first need to integrate the
expression
dP Vd3 ke 4~e2 n()a30 |~e ~ke |2 d3 ke
=
D (fi ) (4.74)
dtdd (2)3
3 m2 c3
(1 + ~ke2 a20 )4
over all solid angles; we had assumed that the electron is photoionised into a free final state, and thus we must have ke2  a2
0 ;
using that, we can approximate
(1 + ke2 a20 )4 a80 ke8 ,

(4.75)

and we introduce a coordinate system in which the photon wave


vector ~k and the polarisation vector ~e are parallel to the z- and xaxes, respectively; then, ~ke is described by the two angles (, ),

sin cos
~ke = ke sin sin ,
(4.76)

cos
and

~e ~ke = ke sin cos ;

integrating over the solid angle then yields


Z 2 Z
Z
4
3
2
d
d sin cos =
d sin3 =
,
3
0
0
0

(4.77)

(4.78)

CHAPTER 4. SYNCHROTRON RADIATION, IONISATION AND RECOMBINATION51


which leads us to write the number of transitions per unit time
and unit photon frequency
dP Vd3 ke 16~e2 n() dke
=
D (fi ) ;
dtd (2)3
32 m2 c3 a50 ke4

(4.79)

finally, if the electron is to become unbound,


~ =
thus
2m
ke =
~

~2 ke2
pe
=
,
2m
2m

!1/2
and

and
~
dke
=
4
ke
2m

!5/2

dke =

(4.80)

md
,
~ke

md
;
~

(4.81)

(4.82)

using
d3 p
2 d
=
,
(2~)3 (2c)3
we finally obtain from (4.73)

(4.83)

!5/2
dP Vd3 ke
16e2 n() ~
d
=
dtd (2)3
32 mc3 a50 2m
2 d
!
= () cn()
(2c)3

(4.84)

and thus, for the cross section,


128e2
~
=
5 7/2 2m
3mca0

!5/2
;

(4.85)

identifying the (dimension-less) fine-structure constant,

e2
,
~c

(4.86)

and Bohrs radius (4.66), we finally write this in the intuitive form
() =
with 0 c/a0 ;

a20

(2)9/2  0 7/2
,
3

(4.87)

Chapter 5
Spectra
5.1

Natural Width of Spectral Lines

5.2

Cross Sections and Oscillator Strengths

the analysis of photo-ionisation is equally applicable to systems


in which transitions occur between two bound levels; as before,
we have from (4.60) and (4.64)
dP 4e2 3fi n(fi )
=
|~xfi |2
3
dt
~c

(5.1)

for the dipole transition probability, where |~xfi | is the matrix element of the position operator ~x between the two bound states;
since the photon flux per unit frequency can be expressed by
the occupation number n() times the number of states in phase
space,
n()c

n()c 4 k2 dk n()2
d3 k
=
=
,
(2)3 d
(2)3 d
22 c2

(5.2)

we can identify the cross section


!
!
4e2 3fi n(fi )
22 c2
2
|~xfi | D ( fi )
() =
~c3
n(fi )2fi
4e2 fi
|~xfi |2 2D ( fi )
~c
e2

ffi () ,
mc

(5.3)

in which
() = 2D ( fi )
52

(5.4)

further reading: Shu, The


Physics of Astrophysics, Vol I:
Radiation, chapters 2223; Rybicki, Lightman, Radiative Processes in Astrophysics, chapters 10; Padmanabhan, Theoretical Astrophysics, Vol. I: Astrophysical Processes, sections
7.17.3

CHAPTER 5. SPECTRA

53

is the line profile function, and ffi is called the oscillator


strength; the factor
e2
(5.5)
mc
has the dimension length2 time1 , the profile function has the dimension time, i.e. the cross section (5.3) does have the dimension
of an area, as it needs be; the profile function is normalised such
that its integral over frequency is unity,
Z
Z
2
D ( fi )d =
D ( fi )d = 1 ;
(5.6)
0

accordingly, the oscillator strength is


ffi =

mc 4mfi
4e2 fi
|~xfi |2 2 =
|~xfi |2 ;
3~c
e
3~

(5.7)

it is dimension-less because its unit is


g s1 cm2 g cm2 s2
= 2
=1;
erg s
s g cm2

(5.8)

for electric dipole transitions, f 1 typically;

5.2.1

Transition Probabilities

transitions between bound states occur spontaneously only with a


certain transition probability, i.e. there is an uncertainty in energy
corresponding to their uncertainty in time; if 1 is the mean life
time of the initial state, the uncertainty in energy is approximately
E

1
~

E ~ ;

(5.9)

therefore, spectral lines are not infinitely sharp but somewhat


widened; the profile function () will thus not typically be a
D function; we shall now calculate the shape of this function if
the mean life time 1 is finite (rather than infinite);
doing so, we consider a radiative transition between two bound
states 1 and 2 with energy eigenvalues E1 and E2 ; generally, the
Schrodinger equation requires for the amplitude an of the energy
eigenstate n
i~

an X
=
hn|H|mi am ei(En Em )t/~ ,
t
m

(5.10)

where H is again the interaction Hamiltonian between the electron and the electromagnetic field;

CHAPTER 5. SPECTRA

54

modeling the transition between the states |2i and |1i, we can set
a2 = 1 and am = 0 for m , 2; due to the finite life time of the state
|2i, we can use the ansatz
a2 (t) = et/2

(5.11)

and obtain
a1
i~
= h1|H|2i a2 (t) ei[(E1 E2 )/~]t
t
)
( "
#
E1 E2 + ~
t
= h1|H|2i exp i
t
~
2
#
"
t
= h1|H|2i exp i( 12 )t
(5.12)
2
where

E1 E2
(5.13)
~
is the angular frequency corresponding to the transition energy;
upon integrating (5.12) with the boundary condition a1 (t = 0) =
0, we find
12

a1 (t) =

ih1|H|2i 1 exp [i( 12 )t t/2]


;
~
12 + i/2

(5.14)

in the limit of very long times, t  1 , thus t  1, part of the


exponential factor can be approximated as
et/2 0 ,

(5.15)

and the transition rate becomes


2

|a1 (t)|

|h1|H|2i|2
=
~2
|h1|H|2i|2
=
~2

12 + i/2
1
;
( 12 )2 + 2 /4

(5.16)

this changes the profile function to read


() =

,
( 12 )2 + (/2)2

(5.17)

which is again defined such that its integral over frequency


(rather than angular frequency ) is unity;
we now need to determine ; we had seen above that the transition
rate is given by (5.1), which can be written as
e2 2fi e2 212 n(12 )
dP 4mfi n(fi )
=
|~xfi |2
=
f12 ,
(5.18)
dt
3~
mc3
mc3
which corresponds to the decay rate; since f12 1 for the transitions which are of interest here, a good estimate for is

e2 212
;
mc3

(5.19)

CHAPTER 5. SPECTRA

55

the cross section for the line transition is


12 () =

e2
e2

; (5.20)
f12 () =
f12
mc
mc
( 12 )2 + (/2)2

in the centre of the line, i.e. at = 12 , we have


e2
4 4e2
f12 2 =
f12 ;
mc

mc
using from (5.18), we find
!2
c
4e2 mc3
= 4
212 ,
12 =
mc e2 212
12
12 =

(5.21)

(5.22)

i.e. the cross section in the centre of the line is proportional to the
square of the absorbed wave length;

5.3

Collisional Broadening of Spectral Lines

collisions between atoms can change the occupation numbers and


the life times of states and modify the line profile of emission or
absorption in this way; the effect of collisions can be described
by random changes of the phase of a1 ,
a1
= h1|H|2i exp [i( 12 )t t/2] ei(t) ,
t
where (t) is a random function such that

1 if there was no collision

D
E

i(t)
until time t
e
=
,

0 else
i~

(5.23)

(5.24)

which means that the average phase factor (5.24) expresses the
probability that the individual system under consideration experienced no collision until time t; in this way, (5.23) formalises the
expectation that (sufficiently energetic) collisions can change the
phase of a1 completely;
extending this consideration from a single system to an ensemble
of systems and averaging over them, the ensemble average will
turn into an exponential if we assume that the number of collisions in the ensemble until time t follows a Poisson distribution,
D
E
ei(t) ec t/2 ,
(5.25)
in which 1
c is the mean time between collisions; using this, we
find the change in a1 after time t to be proportional to the integral
of (5.23),
"
#
Z
t
a1
dt exp i( 12 )t + i(t) ;
(5.26)
2

CHAPTER 5. SPECTRA

56

the averages over time and over all systems in the ensemble then
yield, using (5.25),
"
#
Z
t D i(t) E
dt exp i( 12 )t
e
ha1 i
2
#
"
Z
t c t
; (5.27)
=
dt exp i( 12 )t
2
2
obviously, therefore, the sum of the mean decay and collision
rates and c now takes the role that had before, i.e. the collisions shorten the mean life time to
1
1
1
1

+ c 1 + c /

(5.28)

and thus broadens the line profile;

5.4

Velocity Broadening of Spectral Lines

a further broadening mechanism is caused by the Doppler effect;


if the emitting atoms (or molecules) move along the line-of-sight,
we observe the frequency

vk 
= 0 1 +
(5.29)
c
instead of the frequency 0 , if vk is the velocity component along
the line-of-sight;
it is often appropriate the assume a Gaussian velocity distribution;
the observed line profile is then given by
"
#
Z


dvk
vk 
(vk v)2
D 0 1 +
,
(5.30)
exp
p
c
22v

22v
where v is the mean velocity of the emitting system, and v is the
velocity dispersion of its particles; using the identity
1
D (ax) = D (x) ,
a

(5.31)

the Gaussian line profile

!2
1 0

1
c
exp 2
c v
p
0 22v
2v
0

!
c2 2
c
1

=
exp 2
p
0 22v
2v 0

(5.32)

CHAPTER 5. SPECTRA

57

follows, where we have defined



v 
0 1 +
,
c

(5.33)

i.e. is the central line frequency, shifted by the Doppler effect


due to the mean motion of the emitting or absorbing medium;
according to the definition (5.3) of the scattering cross section,
we have for the Doppler-broadened line

!
c2 2
e2
c
1
;
=
exp 2
f12
(5.34)
p
mc
0 22v
2v 0
if the motion of the particles in the medium is thermal, we obtain
for the velocity dispersion
m 2 kT
=
2 v
2

2v =

kT
,
m

which gives the cross section


r

!
mc2 2
e2
1
mc2
;
=
f12
exp
mc
0 2kT
2kT 0

(5.35)

(5.36)

the cross section in the centre of the line, i.e. at = , is thus


r
mc2
e2 f12
=
(5.37)
mc 0
2kT
and thus proportional to the inverse root of the temperature;

5.5

The Voigt Profile

natural line width and collisional broadening result in the Lorentz


profile, while the (typically thermal) Doppler broadening results
in a Gaussian profile; if all effects need to be taken into account,
the resulting profile is a convolution of a Lorentz and a Gaussian
profile,

Z
v2k

1
() =
exp 2 dvk ,
p
2
2
2v
( 12 ) + (/2)
22v
(5.38)
where we have neglected for simplicity that the line may be
shifted as a whole due to the mean motion with velocity v;
in (5.38), we need to replace 12 by

vk 
12 1 +
c

(5.39)

CHAPTER 5. SPECTRA

58

to take the Doppler shift of the emission frequency into consideration; this yields
Z

() =
2
2
( 12 12 vk /c) + (/2)

v2k
1
exp 2 dvk ;
(5.40)
p
2v
22v
we now set
v0

2v ,

and

12 c
,
12 v0

c
212 v0

vk
vk
q
=
2v v0

(5.41)

(5.42)

and obtain
2ac
() =
v0 12

eq dq
,
(u q)2 + a2

(5.43)

which is the so-called Voigt profile;


near its centre, this line profile has a Gaussian shape, while its
wings retain the Lorentzian shape;

5.6

Equivalent
Growth

Widths

and

Curves-of-

two concepts have been introduced for describing the information


contained in observed spectral lines, namely the equivalent width
and the curve-of-growth;
the equivalent width quantifies the area under a spectral line; if I0
is the specific intensity of the spectral continuum, the equivalent
with is defined as
Z
I0 I()
W
d ,
(5.44)
I0
where I() is the spectral (specific) intensity within the line; thus,
the equivalent width of an absorption line is a measure for the
intensity removed from the spectrum, or added to the spectrum
by an emission line;
the optical depth within the line is
= N L () ,

(5.45)

CHAPTER 5. SPECTRA

59

where N is the number of absorbers and L is the extent of the


absorbing medium; the specific intensity is then
I() = I0 e ,

(5.46)

and thus the equivalent width is


Z
h
i
W=
d 1 e() ;

(5.47)

since the cross section is proportional to the profile function (),


(5.47) can equally be written as
Z
h
i
W=
d 1 eC() ,
(5.48)
with a constant (frequency-independent) C;
for small optical depths,  1, the exponential function in (5.46)
or (5.47) can be expanded into a Taylor series; this results in
Z
e2
W=
d N L () = N L
f12
(5.49)
mc
because the profile function had been normalised such that its integral over frequency yields unity; thus, for small optical depths,
we have
WN,
(5.50)
i.e. the equivalent width is simply growing linearly with the number of absorbers;
if  1, the function
1 e

(5.51)

behaves like a step function across the absorption line whose step
width is determined by how approaches unity; let be this step
width in frequency space, then
W 2 ;

(5.52)

for Doppler-broadened lines, we have


r
"
!#
e2 f12
mc2
mc2
= N L = N L
exp
;
mc 0
2kT
2kT 0

(5.53)

the condition 1 then requires


#
!1
!1/2
"
mc2 2 ! 1 e2 f12
mc2
C1 1
=

(5.54)
exp
2
2kT 0
NL mc0
2kT
C2 NL
with the abbreviations
r
C1

2kT
0 ,
mc2

C2

e2
f12 ;
mc

(5.55)

CHAPTER 5. SPECTRA

60

from this and (5.54), we obtain


!
NLC2
2kT 2
2
ln

=
mc2 0
C1

!
C1 1/2 NLC2
= ln
,
C1

(5.56)

i.e. W(N) grows approximately as ln N;

for a Lorentz profile,


() =

,
( 12 )2 + (/2)2

(5.57)

we can approximate the cross section in the limit | 12 | 


by
e2
C2

=
=
;
(5.58)
f12
2
mc
( 12 )
( 12 )2
again, we conclude from
!

= N L =

NLC2 = 2

i.e. the equivalent width grows in this case as

NLC2 ,
(5.59)

N;

summarising, the curve-of-growth W(N) behaves as

N
small N

1/2
ln
N
intermediate
N
W(N)

N
large N

(5.60)

for determining N, lines with different oscillator strengths f are


used because then the spectral lines fall into different sections of
the curve-of-growth W(N) for the same number N of absorbers;
this may prove difficult when
some lines fall into the flat section
of W(N) where W(N) ln N;

Chapter 6
Energy-Momentum Tensor and
Equations of Motion
6.1

6.1.1

Boltzmann Equation
Momentum Tensor

and

Energy-

Boltzmann Equation

there are many possible approaches to hydrodynamics; one of


the most direct ones starts by considering the particles of a fluid
as they interact by collisions; the distribution of the particles in
phase space (~x, ~p) then becomes the fundamental physical quantity;
let f (~x, ~p, t) be the distribution function in phase space, i.e. the
quantity
dN = f (~x, ~p, t) d3 xd3 p
(6.1)
is the number of particles in the six-dimensional phase space element d3 xd3 p;
forces acting on the particles can be distinguished according to
whether they are smooth or rough on microscopic length
scales; forces which are rough on a microscopic scale are due to
the direct interactions between the particles and are summarised
as collisions; forces which are smooth on a microscopic scale
are described by a potential U,
~ ;
F~ = U

(6.2)

without collisions, the distribution function would satisfy the


collision-less Boltzmann equation,
d f (~x, ~p, t)
=0,
dt
61

(6.3)

further reading: Shu, The


Physics of Astrophysics, Vol II:
Gas Dynamics, chapters 13;
Padmanabhan, Theoretical Astrophysics, Vol. I: Astrophysical Processes, sections 8.18.4;
Landau, Lifshitz, Theoretical
Physics, Vol VI: Hydrodynamics, chapter I

CHAPTER 6. ENERGY-MOMENTUM TENSOR AND EQUATIONS OF MOTION62


in which the time derivative is to be taken along the particle trajectories; this equation says that in absence of collisions, particle
trajectories are not lost in phase space, which is a consequence of
Liouvilles theorem for Hamiltonian systems;
writing the time derivative in (6.3) explicitly, the collision-less
Boltzmann equation reads
f f
f
+ ~x
+ ~p
=0
t
~x
~p

(6.4)

~
or, using (6.2) and ~p = F,
f ~
~ f = 0
+ ~x f U
t
~p

(6.5)

in presence of collisions, the right-hand side of the Boltzmann


equation is changed from zero to
0 C[ f (~x, ~p, t)]

(6.6)

where C[ f ] is a functional of the distribution function f which


describes how it is changed by collisions;
collisions happen between particles with momenta ~p1 and ~p2 and
lead to momenta ~p01 and ~p02 ; the scattering cross section as a function of the solid angle be (); if the distribution functions are
abbreviated as follows
f (~x, ~p1 , t) f1
f (~x, ~p01 , t) f10

,
,

f (~x, ~p2 , t) f2
f (~x, ~p02 , t) f20 ,

the collision term can be written as


Z


C[ f ] =
dd3 p2 () |~v1 ~v2 | f10 f20 f1 f2 ;

(6.7)

(6.8)

this is a result of kinetic theory; we shall later see that this term
drops out when moments of the Boltzmann equation are formed;

6.1.2

Moments; Continuity Equation

hydrodynamics builds upon the central assumption that the mean


free path of the particles making up a fluid is very much smaller
than the extent L of the system under consideration,  L; this
means that integrations over phase space cells typically average
over many collisions;

CHAPTER 6. ENERGY-MOMENTUM TENSOR AND EQUATIONS OF MOTION63


this suggests to solve Boltzmanns equation by taking its moments (as we did before when studying radiation transport); we
~ = 0,
assume for now that there are no external forces, F~ = U
and form the first and second moments of the equation
f ~
+ ~x f = C[ f ] ;
t

(6.9)

the moments are formed by multiplying Eq. (6.9) with 1 or


the four-momentum p and integrating over momentum space,
weighting with a factor E 1 (~p), in short, by applying the operators
Z 3
Z 3
d p
d p
and
p
(6.10)
E(~p)
E(~p)
to (6.9);
because of momentum conservation during collisions, the collision terms on the right-hand side of (6.9) drop out: no net momentum is exchanged on average because under the basic underlying assumption  L, collisions are very frequent and it can be
assumed that collisions away from ~p and to ~p are in equilibrium; note that this is different from starting with a collision-less
equation in the first place; here, the collision terms are present,
but have no net effect because of the very short mean free path;
the first moment yields
Z 3
Z 3
d p f
d p ~
~x f = 0 ;
+
E(~p) t
E(~p)

(6.11)

using the definition


Z

J c
we find
J =
0

d3 p
p f (x , ~p) ,
E(~p)
Z

d3 p f (x , ~p)

(6.12)

(6.13)

because p0 = E/c; quite obviously, J 0 is the particle density n(x )


at the position ~x and the time t = x0 /c;
similarly, the spatial components of J are
Z 3
d p i
i
p f (x , ~p) ,
J =c
E(~p)

(6.14)

or, using the expressions E = mc2 and ~p = m~x which are valid
for relativistic and non-relativistic particles alike,
Z
~x
J~ =
d3 p f (x , ~p) ;
(6.15)
c

CHAPTER 6. ENERGY-MOMENTUM TENSOR AND EQUATIONS OF MOTION64


since the average particle velocity is
R
Z
d3 p ~x f
1
R
~v =
d3 p ~x f ,
=
3
n
d pf

(6.16)

the spatial vector J~ turns out to be the average particle current n~v
divided by the light speed c; therefore, the first moment equation
can be written in the form
J 0 J i n ~
+ i =
+ (n~v) = 0 ,
x0
x
t

(6.17)

upon multiplying this equation for the evolution of the number


density n with the particle mass m, we obtain the continuity equation for the mass,
~
+ (~v) = 0 ,
(6.18)
t
which can also be expressed by the vanishing four-divergence
J
=0
x

(6.19)

of the current J , which is the relativistic generalisation of the


continuity equation (6.18);

6.1.3

Energy-Momentum Tensor

we now consider the second moments, which are obtained by applying the second integral operator from (6.10) to the Boltzmann
equation (6.9); we first study the tensor
Z 3
d p
2
c
p p f T ;
(6.20)
E(~p)
in the non-relativistic limit, we have

~x2
E mc2
p = =
= mc 1 + 2 ,
c
c
2c
0

and thus the time-time component of the tensor T is

~x2
00
2
3
T
= mc
d p 1 + 2 f
2c
Z
m~x2
= c2 + d3 p
f = c2 + n ,
2
where  is the mean kinetic energy of the fluid particles
R
Z
d3 p m2 ~x2 f
1
m
 R
=
d3 p ~x2 f ;
3
n
2
d pf

(6.21)

(6.22)

(6.23)

CHAPTER 6. ENERGY-MOMENTUM TENSOR AND EQUATIONS OF MOTION65


the space-time part of the tensor T reads
Z
Z
3
i
0i
T
= c d pp f =
d3 p mc xi f

!i
Z

~x2
~q
3
i

d p 1 + 2 mc x f = c~v +
, (6.24)
2c
c
where ~q is the current of the kinetic energy
2
Z
m~x
3
~x f ;
~q
d p
2

(6.25)

the first term on the right-hand side of (6.24), c~v, is the mass
current times the light speed;
finally, the space-space components of T are
Z 3
Z 3
d p
d p i j
2
ij
2
pp f =c
T = c
m2 xi x j f
mc2
E(~p)
Z

d3 p m xi x j f ,
(6.26)
which can be interpreted as the (three-dimensional) stress-energy
tensor;
we now return to the (simplified) Boltzmann equation (6.9) and
rewrite it somewhat, using
E = mc2 ,

~p = m~x

~p
~x
= 2
E c

(6.27)

as well as p0 = E/c and x0 = ct; with that, we first obtain


~p ~
f
+ c2
f = C[ f ] ,
0
x
E

(6.28)

~p ~
p0 f
+ c2
f = C[ f ] ;
0
E x
E

(6.29)

c
and from that,
c2

this can obviously be written in the form of a four-divergence,


!
!
0

p
2p
2~
fc
+
= C[ f ] ,
(6.30)
fc
x0
E
E
~x
R
and an integration over d3 p p yields
T 0
=0,
x

(6.31)

i.e. this moment of the Boltzmann equation is equivalent to the


vanishing four-divergence of T 0 ;

CHAPTER 6. ENERGY-MOMENTUM TENSOR AND EQUATIONS OF MOTION66


we can proceed in much the same way with the spatial components T i of T ; doing so, we first write
cpi

f
~ f = cpi f + pi x j f = C[ f ] pi
+ pi ~x
0
x
x0
x j

and insert
x j = c2
to find

pj
E

(6.32)

(6.33)

j
E f
2 i p f
c p
+c p
= C[ f ] pi ,
0
j
cE x
E x

(6.34)

c2
c2 0 i f
f
p p 0 + pi p j j = C[ f ] pi ;
E
x
E
x

(6.35)

2 i

which equals

after integration over d3 p, this yields


i0
i j T i
T
+
T =
=0;
x0
x j
x

(6.36)

thus, the moment equations of the Boltzmann equation can


be expressed by the vanishing four-divergence of the energymomentum tensor T ,
T
=0;
x

(6.37)

this indicates how the hydrodynamical equations can be relativistically generalised; we now return to the non-relativistic expressions for T ;
the time-component of the divergence equation implies
!

~
1  2
q
~ c~v +
c + n +
=0;
c t
c

(6.38)

using the continuity equation (6.18) to eliminate the partial time


derivative of the density from (6.38), we find
 
~ (~v) + n + c
~ (~v) + 1
~ ~q = 0
c
(6.39)
t c
c
and thus

(n ) ~
+ ~q = 0 ;
(6.40)
t
this is the continuity equation for the energy, i.e. the expression
of energy conservation;

CHAPTER 6. ENERGY-MOMENTUM TENSOR AND EQUATIONS OF MOTION67


the spatial part of the divergence equation implies
!
qi

1
i
cv +
+ j Tij = 0 ;
c t
c
x

(6.41)

here, we can neglect the term ~q c2 because the energy flow will
be much slower than the speed of light, and obtain
(vi ) T i j
=0,
+
t
x j

(6.42)

which is the equation of momentum conservation in the fluid approximation;


summarising, we obtain the equations of motion
(n ) ~
+ ~q = 0
t
(vi ) T i j
+
=0
t
x j

,
,

2
m~x
~x f
with ~q
d p
2
Z
ij
with T
d3 p m xi x j f (6.43)
Z

we now split the velocities ~x of the particles into the mean velocity
~v and a (usually thermal) velocity ~u about the mean,
~x = ~v + ~u ;

(6.44)

the kinetic energy then reads


Z
Z
Z
2
m
3 m 2
3 m
~v + ~u f =
n =
d p
d p ~v f + d3 p ~u2 f ,
2
2
2
(6.45)
because the (thermal) velocities ~u vanish on average, by definition; we thus obtain

nm D 2 E
~u ,
(6.46)
n = ~v2 +
2
2
D E
where ~u2 is the mean-squared thermal velocity,
R
D E
d3 p ~u2 f
~u2 R
;
d3 p f

(6.47)

the second term in (6.46) is the internal energy , the first is the
kinetic energy of the mean fluid motion; if the internal energy is
thermal,
3
 = nkT ;
(6.48)
2

CHAPTER 6. ENERGY-MOMENTUM TENSOR AND EQUATIONS OF MOTION68

6.2
6.2.1

The Tensor Virial Theorem


A Corollary

an important theorem for all systems satisfying the equations we


have derived is the tensor virial theorem, which holds independent
of the particular form of the energy-momentum tensor T ; in
order to prove it, we first demonstrate that the relation
Z
Z
d
dF
3
d x F =
d3 x
(6.49)
dt V
dt
V
holds for arbitrary functions F(~x, t) and integration volumina V;
the proof proceeds as follows: first, the total time derivative of the
integral is
!Z
Z
Z
d

(F)
3
3
~
d x F =
+ ~v
d x F =
d3 x
(6.50)
dt V
t
t
V
V
because the integration over d3 x removes the dependence on ~x
and makes the gradient vanish; then, the remaining integral over
the partial time derivative is
!
Z
Z
F
3 (F)
3
~ (~v)
d x
=
d x
F
t
t
V
V
!
Z
F ~
3
~
(F~v) + ~v F
=
d x
t
V
!
Z
F
3
~
=
d x
+ ~v F
t
V
Z
dF
,
=
d3 x
(6.51)
dt
V
where Gauss theorem was employed again to remove the divergence term; this completes the proof;

6.2.2

Second Moment of the Mass Distribution

we now consider the second spatial moment of the mass distribution,


Z
d3 x xi x j

Ii j

(6.52)

and its second time derivative,


" Z
#
Z
d d
d2 I i j
d
d(xi x j )
3
i j
3
=
d
x
x
x
=
d
x

,
dt2
dt dt V
dt V
dt

(6.53)

where the theorem (6.49) was used with F = xi x j ; notice that the
volume V is fixed, so that the coordinates xi introduced in (6.52)
do not explicitly depend on time;

CHAPTER 6. ENERGY-MOMENTUM TENSOR AND EQUATIONS OF MOTION69


the integral on the right-hand side of (6.53) can be transformed to
read
Z
Z
d(xi x j )
3
~ i x j)
d x
=
d3 x ~v (x
dt
V
ZV
(xi x j )
=
d3 x vk
xk
ZV


=
d3 x vk ik x j + vk xi kj
ZV


=
d3 x vi x j + v j xi ;
(6.54)
V

we now take the second total time derivative remaining from


(6.53), which can again be replaced by a partial time derivative
as in (6.50) before,
#
"
Z
d2 I i j
(vi x j ) (v j xi )
3
=
d x
;
+
(6.55)
dt2
t
t
V
now we establish a connection with the energy-momentum tensor
by noting that
jk
j
(xi T jk )
jk i
i T
ji
i (v )
= T k + x
=T x
xk
xk
t

(6.56)

holds because of momentum conservation (6.42); we can use


again that the coordinates xi do not explicitly depend on time to
write this last result in the form
jk
(xi v j )
ji
i T
=T x
t
xk

(6.57)

inserting this into (6.55), we can bring the time derivative of I i j


into the following form:
"
Z

#
d2 I i j
 il j
3
ij
ji
jl i
=
d x T +T l T x +T x ;
(6.58)
dt2
x
V
the second term in this last equation is a divergence, which may be
transformed by Gauss theorem into a vanishing boundary term;
the symmetry of T i j then implies
Z
d2 I i j
= 2 d3 x T i j ,
(6.59)
2
dt
V
which is the tensor virial theorem;
although the described prorcedure of taking moments is mathematically meaningful and correct, it has not brought us closer to a
solution of Boltzmanns equation: we do not know the form of the
stress-energy tensor yet; only when this is defined can we solve
the equations of motion;

CHAPTER 6. ENERGY-MOMENTUM TENSOR AND EQUATIONS OF MOTION70


here, too, there are numerous possibilities for closing the moment equations; they are similar to the procedure which we have
applied to describe radiation transport in the local thermodynamical equilibrium: the fluid approximation asserts that the mean free
path, , of the fluid particles is much smaller than the dimension
of the system; accordingly, we can define a small parameter

,
L

1

(6.60)

and expand the distribution function in powers of , such as


f = f0 + f1 + 2 f2 + . . . ,
where f0 is the distribution function in the limit 0;

(6.61)

Chapter 7
Ideal and Viscous Fluids
7.1
7.1.1

Ideal Fluids
Energy-Momentum Tensor

we first start with a distribution function f0 which describes an


infinitely extended medium in thermal equilibrium; then, f0 is the
Maxwellian distribution
!
m~u2
4n
exp
f0 =
,
(7.1)
2kT
(2kT m)3/2
which contains the velocity ~u relative to the mean velocity of the
fluid, according to our previous notation
~u = ~x ~v ;

(7.2)

the stress-energy tensor is then


Z
ij
T =
d3 p m(vi + ui )(v j + u j ) f0 =
Z
Z
i j
3
= mv v
d p f0 + m d3 ui u j f0
Z
Z
i
3
j
j
+ mv
d p u f0 + mv
d3 p ui f0 ;

(7.3)

the two latter terms vanish because f0 is isotropic, the first term is
Z
i j
mv v
d3 p f0 = vi v j ,
(7.4)
and the second term is
Z
m d3 p ui u j f0 = 0
71

for i , j

(7.5)

further reading: Shu, The


Physics of Astrophysics, Vol II:
Gas Dynamics, chapters 34;
Padmanabhan, Theoretical Astrophysics, Vol. I: Astrophysical Processes, sections 8.58.7;
Landau, Lifshitz, Theoretical
Physics, Vol VI: Hydrodynamics, chapters I and II

CHAPTER 7. IDEAL AND VISCOUS FLUIDS

72

and

nm D 2 E
~u
3
which is the gas pressure P,
P=

for i = j
D 2E
~u ;
3

(7.6)

(7.7)

this interpretation also follows from our earlier considerations;


we had
3

 = nkT = h~u2 i h~u2 i = 3nkT = 3P


(7.8)
2
2
in case of thermal motion;
in this way, we obtain the complete stress-energy tensor,
T i j = vi v j + Pi j ;
in addition, we require the flux ~q of the kinetic energy,
Z


m
~v + ~u 2 ~u + ~v f0 ;
~q =
d3 p
2
we write it in components,
Z

m 2
i
~v + ~u2 + 2~u ~v (vi + ui ) f0
q =
d3 p
2

2 i 2 i
= ~v v + ~v hu i + vi h~u2 i + hui~u2 i
2
2
2
2
+ vi v j hu j i + v j hui u j i ,

(7.9)

(7.10)

(7.11)

where we have used again the averages of arbitrary quantities Q,


Z
1
hQi
d3 p (Q f ) ;
(7.12)
n
due to the isotropy of f0 , the second, fourth and fifth term on the
right-hand side of (7.13) vanish, the third term equals vi , and for
the last term we use again
Z
m d3 p (ui u j f0 ) = Pi j = hui u j i ,
(7.13)
thus
v j hui u j i = v j Pi j = Pvi ;
then, the flux of kinetic energy becomes
!
!
~v2  P
~v2
~q =
+ +
~v
+ w ~v ,
2
2
where

(7.14)

(7.15)

+P
(7.16)

is the heat function (enthalpy) per mass; the enthalpy occurs


here instead of the energy because the the pressure work exerted
by the fluid needs to be taken into account;
w

CHAPTER 7. IDEAL AND VISCOUS FLUIDS

7.1.2

73

Equations of Motion

substituting now the expressions (7.9) for the stress-energy tensor


T i j and (7.15) for the flux of kinetic energy into the equations of
motion, we obtain the equations of motion for an ideal fluid:
first, the equation of continuity,
~

~ +
~ ~v = d +
~ ~v = 0 ; (7.17)
+ (~v) =
+ (~v )
t
t
dt
next, the equation of energy transport,
!
" 2
! #
~
v
~v2
~
+ +
+ w ~v = 0 ,
t 2
2

(7.18)

and finally the equation for the momentum transport,



(vi )

+ j vi v j + Pi j = 0 ;
t
x

(7.19)

we rewrite the last two equations in order to bring them into a


particularly manageable form; from energy conservation (7.18),
we have
~v2 ~v2 
+
+
0 =
2 t 2 t
t
2
2
~
~
v
v
~ +
~ (~v) + (~
~ v) +
~ (P~v) ,
+ ~v
2
2

(7.20)

which we can simplify with the aid of the continuity equation,


2
~v2 
~ ~v + (~
~ v) +
~ (P~v) = 0 ;
+
+ ~v
2 t
t
2

(7.21)

momentum conservation requires


vi

+
t
t
vi
v j P

+ vi v j j + v j j + v j j + j i j ,
x
x
x
x

0 = vi

(7.22)

which can be written in vector form as

~v
+
t
t
~
~ v + ~v(
~ ~v) + P
~ ;
+ ~v(~v ) + (~v )~

0 = ~v

(7.23)

for simplifying that, we first use the continuity equation again in


its form

~ +
~ ~v = 0
+ ~v
(7.24)
t

CHAPTER 7. IDEAL AND VISCOUS FLUIDS

74

to get rid of the first, third, and fifth terms on the right-hand side
of (7.23); we thus obtain

~v
~ v + P
~ =0;
+ (~v )~
t

moreover, we can use the identity


!
~v2
~
~
~ ~v) ~v
(~v )~v =
+ (
2
to find

!
~
~v ~ ~v
~ ~v) ~v = P ;
+
+ (
t
2

(7.25)

(7.26)

(7.27)

multiplying (7.25) with ~v yields


~v2
~ v2 + ~v P
~=0;
+ (~v )~
2 t
2

(7.28)

inserting this into the equation of energy conservation, the first,


second, and next-to-last terms cancel to yield
 ~
~ ~v ;
+ (~v) = P
t

(7.29)

using here the energy per mass, /, this equation reads
() ~

~ (~v) + ~v
~
+ (~v) = + +
t
t
t !

~ = P
~ ~v ;
(7.30)
=
+ ~v
t
we have now arrived at Eulers equations,
~
+ (~v) = 0
t
~
~v
~ v = P
+ (~v )~
t

 ~
~ ~v ,
+ (~v) = P
t

(7.31)

which described the conservation laws for mass, energy, and momentum in the approximation of an ideal fluid;
had we allowed (conservative) external forces, with
~
F~ext = m~v = m

(7.32)

with the potential of the force, the right-hand side of the


momentum-conservation equation had acquired an additional potential gradient,
~
~v
~ v = P
~ ;
+ (~v )~
t

(7.33)

CHAPTER 7. IDEAL AND VISCOUS FLUIDS

7.1.3

75

Entropy

the entropy of an ideal, monatomic gas is


!
3k
P
s=
ln
2m

(7.34)

per unit mass, where we have omitted an additive constant; =


5/3 is the adiabatic index; the total time derivative of this specific
entropy is
!

s
P

+ (~v )s
=
P
(7.35)
t
P
t
t
!
!
P
i
+ v i ln
x

1 P
=

P t t
!

P
i
1

+ v

P i
P
xi
x
i
1 P v P vi
=

;
P t t P xi
xi
according to the continuity equation, we can simplify
"
#
"
#

~
~ (~v)
~ ~v = ~
~ v , (7.36)
+ (~v )
=
+
t
t
and we further had

2
P = nkT =  ,
3

(7.37)

and therefore
1 P 1 
=
P t
 t

and

1 P 1 
=
;
P xi  xi

(7.38)

thus, the entropy equation (7.36) reads


s
1  1
~
~ +
~ ~v
+ (~v )s
=
+ (~v )
t
 "t 
#
1 
~ + 
~ ~v ;
=
+ (~v )
 t

(7.39)

finally, we use
!
5 3
5
3
 = nkT = nkT =
+ 1 nkT =  + P ,
3 2
2
2
which allows us to conclude
"
#
s
1

~ =
~ (~v) + P
~ ~v = 0 ,
+ (~v )s
+
t
 t

(7.40)

(7.41)

because the expression in square brackets vanishes due to energy


conservation;

CHAPTER 7. IDEAL AND VISCOUS FLUIDS

76

we thus obtain the conservation of entropy,


s
~ = ds = 0 ,
+ (~v )s
t
dt

(7.42)

which is intuitively expected in the absence of dissipation;

7.2
7.2.1

Viscous Fluids
Stress-Energy Tensor; Viscosity and Heat Conductivity

so far, we have neglected gradients in the temperature because


we have assumed a Maxwellian velocity distribution belonging
to a fixed temperature for the particles; likewise, the stress-energy
tensor of the ideal fluid does not contain velocity- or density gradients; such terms will now be included;
the form of the corresponding expressions in the stress-energy
tensor and in the energy current ~q can be computed by expanding
the phase-space distribution function to the next order beyond the
ideal-fluid term f0 ; we abbreviate here and justify the form of the
appearing terms by physical arguments;
differential velocity terms can be expressed by the tensor
vij

vi
,
x j

(7.43)

whose trace is the divergence of ~v,


trvij =

vi ~
= ~v ;
xi

(7.44)

we subtract this trace from vij in order to obtain a trace-free residual,


1 ~
vij ij
~v ;
(7.45)
3
this expression describes pure shear flows which deform the
~ ~v describes the commedium, while the part proportional to
pression of the medium; we thus obtain the shear tensor
!
1 i~
i
i
~ ~v ,
j 2 v j j ~v ij
(7.46)
3
in which and are constants which remain to be determined
and describe the strength of shear flows and compression, respectively;

CHAPTER 7. IDEAL AND VISCOUS FLUIDS

77

the tensor vij from (7.43) can be split into a symmetric and an
antisymmetric part:
!
!
1 vi v j
1 vi v j
vi
=
+
+

;
(7.47)
x j 2 x j xi
2 x j xi
if the velocity field is caused by rigid rotation,
~v =
~ ~x ,

vi =  ijk j xk ,

the antisymmetric part turns into

!
j
1 vi v j
1 (kli k xl ) (kl k xl )

2 x j xi
2 x j
xi

1 i
=
k j kij k
2
=  ijk k , 0 ,

(7.48)

(7.49)

while the symmetric part vanishes;


in order to prevent rigid rotation from causing dissipation, we use
only the symmetric part and write from now on
!
1 vi v j
i
;
(7.50)
+
vj
2 x j xi
we augment the energy-momentum tensor of the ideal fluid now
by this shear tensor and obtain
T ij = vi v j + Pij ij ,

(7.51)

in which the minus sign is conventional;


accordingly, we can modify the energy current; first, a temperature gradient will cause an energy current against the gradient
which will be quantified by a heat conductivity ,

T
;
xi

(7.52)

moreover, we need to add a contribution to the energy transport


which is due to the flow of the velocity gradient,
v j ij ;
and therefore the energy current then reads


T
i
2
q = ~v + w vi
v j ij ;
2
xi

(7.53)

(7.54)

CHAPTER 7. IDEAL AND VISCOUS FLUIDS

78

the additional contributions are thus characterised by three coefficients, i.e. the heat conductivity , and the two viscosity coefficients and ;
the form of these coefficients can be computed by inserting the
ansatz f = f0 + f1 for the phase-space distribution function into
the Boltzmann equation and iteratively searching for solutions,
evaluating the collision terms;
must have the dimension (energy density velocity)/(temperature gradient), hence
[] =

erg cm cm
erg
=
;
cm3 s K
cm s K

(7.55)

similarly, one finds that the dimension of is (energy density)/(velocity gradient), or


[] =

7.2.2

erg s
erg s
=
;
3
cm cm
cm3

(7.56)

Estimates for Heat Conductivity and Viscosity

we consider a gaseous system in thermal equilibrium with a temperature T whose particles are moving randomly in all directions;
let A be the area of a screen perpendicular to the y axis;
per unit time,

n v A
(7.57)
6
particles will fly through the screen, either from left to right or
the other way round; the factor of 1/6 is owed to the fact that
typically only 1/3 of the particles is flying along the y axis, and
of those, only 1/2 in either direction;

the mean free path is = (n)1 , if is the collisional cross


section of the particles; particles coming from the left transport
properties of the gas from y to y, and particles coming from the
opposite direction transport properties from y + to y; interesting
effects occur if these properties have gradients;
if the particle number density, n, changes along y, n/y , 0, the
net number of flowing particles is
n(y + )vA n(y )vA vA n

2 ,
6
6
6 y

(7.58)

where we have implicitly assumed that the mean free path is


very short compared to the typical length scale of the numberdensity gradient;

CHAPTER 7. IDEAL AND VISCOUS FLUIDS

79

the diffusion coefficient D, which relates the particle current per


unit area to the number-density gradient,
N
n
=D
,
(7.59)
A
y
thus is
v
;
(7.60)
D=
3
if the temperature changes along y, T/y , 0, the particles transport energy,

nvA 

(y + ) (y )
=
(7.61)
A
6A
!
nv  T
nv cv T
=
=
,
3 T y
3 y
where cv is the heat capacity at constant volume; using
 ! T
=
,
A
y
we find
nv cv v cv
=
=
;
3
3
since  is the energy per particle, we have
3k
cv =
2
and thus
vk
,
=
2
which has the physical unit
cm erg 1
erg
[] =
=
,
2
s K cm
cm s K
as expected from (7.55);

(7.62)

(7.63)

(7.64)
(7.65)

(7.66)

in complete analogy, the transport of momentum is


p x nvA
v x
=
2m
,
(7.67)
t
6
y
i.e. momentum in the x direction is transported in this way along
the y axis;
the change of momentum per unit time is a force; the force per
unit area is
p x
nvm v x ! v x
Fx =
=
=
,
(7.68)
t A
3 y
y
where is the viscosity coefficient; from this, we obtain
nvm mv 2m
=
=
=
,
(7.69)
3
3
3k
which clarifies where the form of the expressions for and originate from;

CHAPTER 7. IDEAL AND VISCOUS FLUIDS

7.2.3

80

Equations of Motion for Viscous Fluids

again, we can find the equations of motion by inserting ~q and T i j


into the general results (6.18) and (6.43), where
3

n = ~v2 + nkT ~v2 +  ;


2
2
2

(7.70)

obviously, the continuity equation remains valid without change;


the force equation now reads, with the stress-energy tensor (7.51)
and the shear tensor (7.46)
(vi ) (vi v j ) P
+ i
+
t
x j
x

~
~ ~v)
vi j 1 (
+ ( ~v) ;
= 2 j
x
3 xi
xi

(7.71)

the right-rand side of this equation can be simplified to read


!
2 v j
v j
vi v j
j
+

x x j xi
3 xi x j
xi x j
 (

~ ~v)
2 i
~
(7.72)
+
;
= v +
3
xi
the left-hand side of (7.71) can be transformed by means of the
continuity equation,

~v
~ + (~v )~
~ v + ~v (
~ ~v) =
+ ~v + ~v (~v )
t
t
!
~v

~
~
~ v=
= + ~v
+ (~v ) + ~v + (~v )~
t
t
d~v
= ;
(7.73)
dt

this leads us to the Navier-Stokes equation,


 ~
d~v
~ +
~ 2 vi + + ( ~v) ,
= P
dt
3
xi

(7.74)

which simplifies to Eulers equation if the viscosity parameters


vanish, = 0 = ;
the energy-conservation equation is
!
"
!
#
~v2
~v2

T
i
ij
+ + i
+w v
v j = 0 ;
t 2
x
2
xi
this expression can be simplified as follows:

(7.75)

CHAPTER 7. IDEAL AND VISCOUS FLUIDS

81

the equation of momentum conservation can be written as




(vi )
+ j vi v j + Pi j i j = 0 ;
t
x

(7.76)

multiplying this with vi and using the continuity equation enables


us to write
!
 2
1 2 j
i j
P
~v + j ~v v = vi
+ vi j ;
(7.77)
t 2
x 2
xi
x
subtracting this from the energy conservation equation yields
!
 (wvi )

T
~ i j v j = 0 ; (7.78)

(~
v

)P
t
xi
xi xi
xi
using the definition of the enthalpy (7.16), we can transform
(wvi ) [( + P)vi ]
~ + P) + ( + P)
~ ~v , (7.79)
=
= (~v )(
xi
xi
and the energy conservation equation can be cast into the form
 ~
~ ~v =
~ (T
~ ) + i j vi ;
+ (~v) + P
t
x j

7.2.4

(7.80)

Entropy

we now introduce again the energy per unit mass, /, to


write
"
#
!
() ~
~

~
+ (~v) =
+ (~v) +
+ ~v
t
t
t
d
(7.81)
= ;
dt
this first implies the equation

d
~ ~v =
~ (T
~ ) + i j vi ;
+ P
dt
x j

(7.82)

~ ~v can also be rewritten; because of the continuity


the term P
equation, we first obtain
d
~ = +
~ (~v)
~ ~v =
~ ~v ,
=
+ ~v
dt
t
t
which allows us to write
!
1 d
P d
d1
dV
~
P ~v = P
=
= P
= P
,
dt
dt
dt
dt
where we have used the specific volume V 1 ;

(7.83)

(7.84)

CHAPTER 7. IDEAL AND VISCOUS FLUIDS

82

thus, the left-hand side of the energy equation (7.82) can be cast
into the form
!
ds
d
d
dV
~ ~v =
= T
+ P
+P
dt
dt
dt
dt
"
#
s
~
(7.85)
+ (~v )s
,
= T
t
where s is again the specific entropy; therefore, we finally obtain
"
#
ds
s
~
~ (T
~ ) + i j vi ;
(7.86)
T
= T
+ (~v )s =
dt
t
x j
this describes how the entropy is changed due to heat conduction
and viscous dissipation; obviously, the entropy is conserved if
= 0 = i j ;

7.3
7.3.1

Generalisations
Additional External Forces; Gravity

the equations derievd so far can be generalised in obvious ways if


~ or a radiation pressure
external forces act such as gravity, ,
~
force frad ; such additional forces leave the continuity equation unchanged; the force equation acquires the additional force density
terms
~ + f~rad ;

(7.87)
in the energy-conservation equation, two additional terms appear
which describe the work done by the external forces, thus
~ + ~v f~rad ;
~v
the stress-energy tensor of the gravitational field is
!
1 1 i j
ij
T =

4G xi x j 2 xk xk

(7.88)

(7.89)

whose trace is
T ii

!
1
1 3

=
;
=
i
k
4G x xi 2 xk x
8G xi xi

(7.90)

the spatial integral of the trace,


Z
Z
1

3
i
d3 x i
=
d x Ti =
8G V
x xi
V
Z
h
i
1
~ ()
~
~ 2
=
d3 x
8G V
Z
1
=
d3 x ,
(7.91)
2 V

CHAPTER 7. IDEAL AND VISCOUS FLUIDS

83

is the potential energy in the gravitational field; in the last equality


of (7.90), we have used the Poisson equation,
~ 2 = 4G ,

(7.92)

and dropped the boundary term which results from the divergence
~ ()
~ when we use Gauss theorem;

then, the total stress-energy tensor of a self-gravitating fluid is


ij
ij
ij
T ges
= T gas
+ T grav
= vi v j + Pi j +
!
1 1 i j
+

4G xi x j 2 xk xk

(7.93)

with the trace

~ 2
()
= ~v + 3P
,
8G
whose volume integral is
!
Z
Z
1
3
i
3
2
d x T i ges =
d x ~v + 3P +
2
V
V
Z
= 2T + U + 3 d3 x P ,
T ii ges

(7.94)

(7.95)

where T and U are the kinetic and potential energies, respectively;


the tensor virial theorem (6.59) tells us
Z
Z
d2 Iii
d2 I i j
3
ij
= 2 d xT

= 2 d3 x T ii ;
2t
dt2
d
V
V

(7.96)

which means that the integral over the trace of T i j must vanish if
the system under consideration is static;

7.3.2

Example: Cloud in Pressure Equilibrium

now, we briefly consider two astrophysical consequences of this


result; first, let a homogeneous, spherical cloud be given of mass
M and radius R which has the temperature T ; it be embedded into
the constant pressure P;
its kinetic energy is
3M
kT ,
2m

(7.97)

and the potential energy is


3
GM 2
with =
(7.98)
R
10
for the homogeneous sphere; for other mass distributions, will
be different but remains of order unity;
U =

CHAPTER 7. IDEAL AND VISCOUS FLUIDS

84

our earlier result now asserts


3kT

M
GM 2

+ 3PV = 0
m
R

(7.99)

for a static configuration; with V = 4R3 /3, we find for the pressure
!
1 GM 2 3kT M

;
(7.100)
P=
4
R4
mR3
the external pressure must thus be reduced by the amount of the
gravitational force compared to the thermal pressure NkT/V of
the gas in the sphere;
at the critical mass
Mcr =

3kT R4
3kT R

=
,
3
mR G
mG

(7.101)

the pressure P vanishes, such that the sphere is in equilibrium


with its self-gravity;

7.3.3

Example: Self-Gravitating Gas Sphere

a further example concerns isolated systems in which the kinetic


energy of the gas is
3
T = ( 1)Uint ,
2

(7.102)

where Uint is the internal energy of the gas and the adiabatic
index; for such a static system, the tensor virial theorem requires
3( 1)Uint + Ugrav = 0 ,

(7.103)

if we denote the gravitational potential energy by Ugrav


the total energy is E = Uint + Ugrav , and thus (7.103) implies


3( 1)E 3( 1) 1 Ugrav
= 3( 1)E (3 4)Ugrav = 0
and therefore
E=

3 4
Ugrav ;
3( 1)

(7.104)

(7.105)

since Ugrav < 0 and E < 0 for a bound system, we require


>

4
;
3

(7.106)

CHAPTER 7. IDEAL AND VISCOUS FLUIDS

85

in order to see what happens in the limiting case 4/3, we


write
1 d2 I
= 3( 1)E (3 4)Ugrav
2 dt2
4/3
1
d2 I
E =
;
Ugrav +
1
6( 1) dt2

(7.107)

for 4/3, the first term on the right-hand side vanishes, and
because of E < 0, we must have
d2 I
<0,
dt2
which typically implies a collapse;

(7.108)

Chapter 8
Flows of Ideal and Viscous
Fluids
8.1
8.1.1

Flows of Ideal Fluids


Vorticity and Kelvins Circulation Theorem

ideal fluids are such in which dissipative effects are unimportant


or absent, i.e. fluids for which we can put the viscosity coefficients
and the thermal conductivity to zero, = = = 0; in such
fluids, the entropy is conserved along flow lines,
ds s
~ =0;
=
+ ~v s
dt
t

(8.1)

Eulers equation for the force per unit mass is


~
~v
~ v = P
~ ;
+ (~v )~
t

(8.2)

here, we employ the identity


~ a + ~a (
~ ~b) + ~b (
~ ~a) (8.3)
~ a ~b) = (~a )
~ ~b + (~b )~
(~
and put ~a = ~v = ~b; then, the relation
~ v2 ) = 2(~v )~
~ v + 2~v (
~ ~v)
(~

(8.4)

follows, and thus


~ v = 1 (~
~ v2 ) ~v (
~ ~v) ,
(~v )~
2

(8.5)

from which we obtain


~
~v
~ ~v) = 1 (~
~ v2 ) P
~ ;
~v (
t
2

86

(8.6)

further reading: Shu, The


Physics of Astrophysics, Vol II:
Gas Dynamics, chapters 6 and
14; Padmanabhan, Theoretical
Astrophysics, Vol. I: Astrophysical Processes, sections 8.6
8.9; Landau, Lifshitz, Theoretical Physics, Vol VI: Hydrodynamics, chapters I, II and VIII

CHAPTER 8. FLOWS OF IDEAL AND VISCOUS FLUIDS

87

the curl of the velocity,


~ ~v ,
~

(8.7)

is called the vorticity of the flow; if we take the curl of Eulers


equation in its form (8.6), we find the equation for the vorticity

~
~

~ P
~ (~v )
~
=

t
~
~
~ (~v )
~ + P ;
=
2

(8.8)

if the pressure P is a function only of and not of other quantities


such as ~v, the gradients of P and must align,
~ k
~
P

~
~ =0;
P

(8.9)

for such barotropic fluids, the vorticity equation simplifies to


~
~ ;
= (~v )
t

(8.10)

we consider now the so-called circulation, which is the line integral over the velocity along closed curves,
I
~v d~l ;
(8.11)
C

we are interested in the total change with time of the circulation


embedded into the flow, i.e. taking into consideration that the contour C is deformed by the flow; we first use Stokes theorem to
write
Z
Z
~
~ dA
~
~,
= ( ~v) dA =

(8.12)
A

~ is the directed
where A is the area enclosed by the contour C; dA
area element pointing along the local normal to the area A;
the total time derivative of now is
Z
Z
~
d
~
~ A ,
=
dA +
dt
t
A t
A

(8.13)

and the change of the area due to the deformation of the contour
is
I
~ (~v d~l) ,

(8.14)
C

for (~vdt) d~l gives the differential change of area per time interval
dt;

CHAPTER 8. FLOWS OF IDEAL AND VISCOUS FLUIDS


using (8.14) yields with (8.13)
Z
Z
d
~
~ ~v) d~l
=
dA + (
dt
t
A
C

Z ~

~
~ ~v) dA
=
+ (

~=0,
t

88

(8.15)

since the term in square brackets vanishes because of the vorticity


equation (8.10); we have further used that
~ (~v d~l) = (
~ ~v) d~l and

~ (~v )
~ (
~ ~v) =
~ ;

(8.16)

thus, the circulation along contours comoving with the flow is


conserved in barotropic flows; this is Kelvins circulation theorem;

8.1.2

Bernoullis Constant

if the flow is stationary, all partial derivatives with respect to time


vanish; in such cases, flow lines can be introduced which are the
integral curves of the velocity field; obviously,
dy dz
dx
= dt =
=
vx
vy
vz

(8.17)

for the flow lines;


in ideal fluids, the specific entropy is constant,
ds s
~ =0
=
+ (~v )s
dt
t

thus

~ =0
(~v )s

(8.18)

for a stationary flow because s/t = 0, i.e. the entropy remains


constant along flow lines; moreover, the enthalpy satisfes
dw = d + Pd(1 ) + 1 dP ,

(8.19)

because it is, per definition,


w=

+P
P
=+ ,

(8.20)

and with
d = T ds PdV = T ds Pd(1 )

(8.21)

dw = T ds + 1 dP = 1 dP ,

(8.22)

we find
since ds = 0 along flow lines;

CHAPTER 8. FLOWS OF IDEAL AND VISCOUS FLUIDS

89

again for stationary flows, ~v/t = 0, Eulers equation in its form


(8.6) implies
~
1~ 2
~ ;
~ = P
(~v ) ~v
2

(8.23)

let now ~l be a unit tangent vector to a flow line; if we multiply


(8.23) with ~l, the first term quantifies the change of ~v2 /2 along the
flow line; the second term vanishes, because it is perpendicular to
~v and thus also to ~l, and this implies that
B

~v2
+w+
2

(8.24)

is constant along flow lines; this is Bernoullis equation,


B = constant along flow lines ;

(8.25)

we have merely used that ds = 0, i.e. such flows can be called


adiabatic because no heat is exchanged between flow lines; if the
flow is furthermore isentropic, i.e. if s = const. holds everywhere
and thus all flow lines have the same value of s, then
s = s(P, ) = const.

(8.26)

means that P must be a function of alone, i.e. the flow is then


also barotropic, and the circulation is conserved;

8.1.3

Hydrostatic Equlibrium

in the hydrostatic case, ~v = 0, and Eulers equation simplifies to


~ =
~ ,
P

(8.27)

~ 2 = 4G

(8.28)

and Poissons equation

relates the gravitational potential to the density; taking the divergence of (8.27) yields

~
~ P = 4G ;
(8.29)


only assumptions on mechanical, but not on thermodynamical
equilibrium entered here; the curl of (8.27) shows that
~ ()
~ =
~
~ ,
0=

(8.30)

which shows that the gradients of and are then parallel to each
other, i.e. and have the same iso-surfaces, i.e. equipotential
surfaces are then also iso-density contours;

CHAPTER 8. FLOWS OF IDEAL AND VISCOUS FLUIDS

90

in the spherically symmetric limiting case, we finally have


~ = dP ~er ,
P
dr
and thus

2
~ f~ = 1 d(r fr )

r2 dr

with

fr f~ ~er ,

!
1 d r2 dP
= 4G ;
r2 dr dr

(8.31)

(8.32)

if there is a barotropic relation, P = P(), this is an ordinary,


second-order differential equation for the density which can directly be integrated;
Eulers equation in its form (8.32) is often applied to selfgravitating systems such as galaxy clusters where the gravitational potential is largely caused by the dark matter, which requires us to separate the gas density gas from the dark-matter
density DM ,
!
1 d r2 dP
= 4GDM ;
(8.33)
r2 dr gas dr
with the equation of state for an ideal gas,
P = nkT =

gas
kT ,
m

where m is the (mean) mass of a gas particle, we find


Z r
r2 k d(gas kT )
= 4G
r02 dr0 DM = GM(r) ,
gas m
dr
0

(8.34)

(8.35)

where M(r) is the dark mass enclosed in a sphere of radius r; thus,


!
dgas
dT
kr2
T
+ gas
=
M(r) =
mGgas
dr
dr
!
r2 kT d ln gas d ln T
=
+
=
mG
dr
dr
!
rkT d ln gas d ln T
+
,
(8.36)
=
mG d ln r
d ln r
i.e. if the two logarithmic gradients can be estimated or determined e.g. from X-ray observations, the dark mass can be found;

8.1.4

Curl-Free and Incompressible Flows

~ ~v = 0 = ,
~ the velocity field
if the velocity field is curl-free,
~ and
can be written as the gradient of a velocity potential, ~v = ,
Eulers equation (8.6) then reads
~
~
()
1~ 2
P
~ ,
+ ~
v =

t
2

(8.37)

CHAPTER 8. FLOWS OF IDEAL AND VISCOUS FLUIDS


or, if we combine the gradients (cf. 8.22)
!
2
~

v
~

+w+ + =0;
t
2

91

(8.38)

the quantity

~v2

+w+ +
t
2
can then only be a function of time,
B0

B0 = f (t) ,

(8.39)

(8.40)

which can be set to zero because it can be absorbed into the ve~ thus,
locity potential without changing the relation ~v = ;
~v2

B =
+w+ +=0,
t
2
0

(8.41)

which generalises Bernoullis equation for such cases in which


~
the velocity field is curl-free, ~v = ,
instead of stationary,
~v/t = 0;
finally, if = const. in addition, like for incompressible fluids, its
~ = 0, and the vorticity equation (8.8) implies
gradient vanishes,
h
i
~ ~v)
(
~ ~v (
~ ~v) ,
=
t

(8.42)

i.e. the flow is then described solely by one equation for the ve~ ~v = 0;
locity field, because the continuity equation shrinks to
~ ~v = 0, i.e. if the flow is
if the velocity field is also curl-free,
~ ~v = 0 and ~v =
~ imply that the
incompressible and curl-free,
velocity potential has to satisfy the Laplace equation,
~ 2 = 0 ;

8.2
8.2.1

(8.43)

Flows of Viscous Fluids


Vorticity; Incompressible Flows

Eulers equation for viscous fluids was




~
~v
~ v = P +
~ 2~v + 1 + (
~
~ ~v) ;
+ (~v )~
t

(8.44)

taking the curl of this equation and using the vector identity
h
i
h
i
~ (~v )~
~ v =
~ ~v (
~ ~v) ,

(8.45)

CHAPTER 8. FLOWS OF IDEAL AND VISCOUS FLUIDS

92

we obtain the equation for the vorticity in viscous fluids,


~

~ (~v )
~ 2
~ =
~




~
 ~ ~

2
~
~
P ~v + ( ~v) ;
+
2
3

(8.46)

here, the first term on the right-hand side describes the diffusion
of vorticity caused by the viscosity , and the second term arises
exclusively through the density gradient;
thus, if the flow is incompressible, = const., the second term
on the right-hand side of (8.47) drops out, and the equation is
reduced to
~

~ (~v )
~ 2
~ =
~ ;

(8.47)
t

in incompressible fluids, the divergence of ~v must also vanish,


~ ~v = 0, thus

~ (~v )
~ v (~v )
~
~ = (
~ )~
~ ,

(8.48)

which allows us to write the vorticity equation as


~

~ v + (~v )
~
~ 2
~ )~
~ =
~ ,
(
t

(8.49)

~
or, if we identify the total time derivative of ,
~
d
~ v =
~ 2
~ )~
~ ;
(
dt

(8.50)

~ ~v = 0 for
the divergence of Eulers equation implies, with
incompressible fluids,
h
i
~2
~ (~v )~
~ v = P ;

(8.51)

~ v = 0, these equations (8.50) and (8.51) determine


together with ~
the flow of incompressible, viscous fluids: is constant, the curl
~ ~v = ,
~ and the divergence of ~v vanishes;
of ~v is the vorticity,
this determines the velocity field, and the pressure P follows from
(8.51);

8.2.2

The Reynolds Number

the only dimensional physical parameter in equation (8.50) for


the vorticity of a viscous, incompressible flow is

,
(8.52)

CHAPTER 8. FLOWS OF IDEAL AND VISCOUS FLUIDS

93

whose dimension is
erg s cm3 cm2
=
,
cm3 g
s

(8.53)

thus squared length over time; a body characterised by a geometrical dimension L moving with velocity u through a viscous fluid
thus introduces the length scale L and the time scale L/u; from
them and , we can form the following dimension-less quantity:
L2 1
uL uL
R=
=
,
L/u

(8.54)

which is the so-called Reynolds number;


since no other dimensional physical quantities occur, the velocity ~v of the fluid can be scaled with the velocity u of the body
and the spatial coordinate ~x with the length scale L; in this way,
dimension-less quantities emerge whose change is described by
quantities which otherwise contain only the Reynolds number;
bodies of different size but otherwise equal shape embedded into
flows which are scaled as described here thus create self-similar
flows as long as the Reynolds number remains the same; conversely, the Reynolds number classifies such self-similar solutions of the flow equations; the transition to ideal fluids is characterised by R ;

8.3
8.3.1

Sound Waves in Ideal Fluids


Linear Perturbations

we now consider small perturbations of a fluid which is otherwise


flowing according to a background solution characterised by density 0 and pressure P0 ; we further transform into a coordinate
frame in which the unperturbed fluid is locally at rest, thus ~v0 = 0;
let the flow be ideal for now; the perturbations be small, and
we can consequently linearise Eulers equation around the background solution, i.e. we write ~v = 0 + ~v0 , P = P0 + P0 , = 0 + 0
and neglect terms of higher than first order in ~v0 , P0 and 0 ; this
transforms the continuity equation to
~
(0 + 0 ) ~
+ (~v) = 0 =
+ (0~v0 ) ;
t
t

(8.55)

the background density 0 must also satisfy the continuity equation, which reads
0
=0
(8.56)
t

CHAPTER 8. FLOWS OF IDEAL AND VISCOUS FLUIDS

94

because we are in a frame locally co-moving with the fluid, ~v0 =


0, and thus the linearised continuity equation reads
0 ~
+ (0~v0 ) = 0 ;
t

(8.57)

the linearised Euler equation reads


~ 0
~v0 P
+
=0
t
0

(8.58)

if terms of higher than first order in the perturbations are ignored


again;
a relation between density and pressure is established as follows:
since the ideal fluid flows adiabatically, we can set
!
P
0
P =
0
(8.59)
s
for the pressure perturbations, where the subscript s denotes that
the derivative must be taken at constant entropy; if we further
neglect the density gradient of the background solution on the
~ 0 = 0, the continuity equation
length scale of the perturbation,
requires
!
P0
P ~ 0
~v = 0 ;
(8.60)
+ 0
t
s
this equation connects the curl-free part of ~v0 to the evolution of
~ we first have
the pressure; if we set ~v0 = ,
!
P0
P ~ 2
=0,
(8.61)
+ 0
t
s
and the linearised Euler equation further requires
~ 0
~ + P = 0 ,

t
0

(8.62)

which means that, up to an irrelevant constant, the pressure perturbation is

P0 = 0
;
(8.63)
t
we thus find the wave equation
!
2
P ~ 2

=0;
(8.64)
t2
s

CHAPTER 8. FLOWS OF IDEAL AND VISCOUS FLUIDS

8.3.2

95

Sound Speed

this is the usual dAlembert equation, whose solutions are arbitrary functions f (x) which propagate with a velocity cs in either
direction, f (x cs t), where
"
! #1/2
P
cs
(8.65)
s
is the sound speed; obviously,
2 f (x cs t) [cs f 0 (x cs t)]
=
= c2s f 00 (x cs t)
2
t
t

(8.66)

and

2 f (x cs t)
= f 00 (x cs t) ,
(8.67)
2
x
if the primes denote derivatives of f (x) with respect to its argument; these solutions represent arbitrarily shaped waves propagating at the sound speed cs into the positive or negative x direction;
the condition imposed during the derivation of this wave equation
was that the perturbations are small; since
v0x =
as well as
P0 = 0

= f 0 (x cs t)
x

(8.68)

= 0 f 0 (x cs t)
t

(8.69)

P0 = c2s 0 ,

(8.70)

and
we find

P0
cs 0
=
;
(8.71)
0 cs
0
thus, this condition is satisfied as long as |v|  cs , i.e. for subsonic flows;
v0x =

sound waves describe how small perturbations propagate through


~ they oscillate in propagation directhe fluid; because of ~v0 = ,
tion, i.e. they are longitudinal waves; however, we have to take
into account that entropy is conserved in an ideal fluid,
ds
=0
(8.72)
dt
along flow lines, which means that entropy perturbations can only
propagate with the flow velocity ~v since they have to remain constant along the flow lines; the same holds for the circulation of
barotropic fluids;
in viscous fluids, sound waves are dissipated, i.e. they are converted to heat;

CHAPTER 8. FLOWS OF IDEAL AND VISCOUS FLUIDS

8.4

96

Supersonic Flows

8.4.1

Machs Cone; the Laval Nozzle

perturbations propagate with the sound speed into all directions


relative to the fluid, i.e. relative to a coordinate frame which is
co-moving with the fluid; within a given time t, they reach all
points around their origin with radius cs t;
if the fluid is flowing with velocity ~v, it depends on the modulus
|~v| where perturbations may propagate to; the situation for |~v| < cs
is illustrated in a diagram (to be inserted);
sound waves can still propagate into all directions, but the situation changes according to another diagram (to be inserted) if
|~v| > cs ; then, as seen from the laboratory frame, sound waves can
only reach points within a cone with the half opening angle
= arcsin

cs
;
v

(8.73)

this means that sound waves cannot reach an area in the direction
of the flow because they are passed by the flow; from the point
of view of a body which is at rest in the laboratory frame, this
implies that the flow is meeting with the body blindly, without
having been informed about its presence by sound waves; this
has far-reaching implications;
as an example for the (steady) transition from sub- to supersonic
flow, we consider a nozzle with variable circular cross section A;
mass conservation requires
vA = const.

(8.74)

or

d(v)
dA
=
;
v
A
from Eulers equation, we obtain in the stationary case
~v

~
~v P
d~v
1 dP
=
=
,
dt

dt

(8.75)

(8.76)

for P/t = 0 due to the assumed stationarity; in the rotationallysymmetric and thus effectively one-dimensional case considered,
this implies
dv
1 dP
=

dt
dt
!
c2
1 P
dP
vdv =
=
d = s d ,

(8.77)

CHAPTER 8. FLOWS OF IDEAL AND VISCOUS FLUIDS


and thus

d
vdv
= 2 ;

cs

97

(8.78)

with this result, we obtain


!
dA
vd + dv
d dv
v2 dv
=
=
= 1 2
A
v

v
cs v
or

!
c2s
d
dA
1 2 ;
=
A

(8.79)

(8.80)

we first consider a flow which enters with v < cs into a nozzle


which widens in the flow direction; in this case, dA/A > 0 and
!
c2s
1 2 >0,
(8.81)
v
thus dv/v < 0; therefore, the flow decelerates, but becomes
denser, d/ > 0;
the reverse case happens if the nozzle narrows in the flow direction, dA/A < 0; then, dv/v > 0 and the flow accelerates;
exactly the opposite occurs if the flow is supersonic on entrance,
v > cs , since then
dA
>0
A

dv
>0
v

(8.82)

and

dA
dv
<0
<0,
(8.83)
A
v
i.e. a supersonic flow accelerates in a widening nozzle; we now
consider a nozzle which is assembled as follows (graphic to be
inserted); if sub-sonic gas is entering the nozzle, v < cs , it accelerates; if it remains subsonic up to the smallest cross section,
its velocity decreases again; however, if it reaches v = cs at the
narrowest cross section of the nozzle, the gas accelerates further
beyond the sound speed;

8.4.2

Spherical Accretion

now we consider the example of spherical accretion, i.e. of a body


of mass M with radius R which is embedded in a gas cloud from
which it attracts mass; the continuity equation requires in this
spherically-symmetric case
1 2
(r v) = 0
r2 r

(8.84)

CHAPTER 8. FLOWS OF IDEAL AND VISCOUS FLUIDS

98

if the accretion flow is stationary and thus /t = 0; the quantity r2 v is therefore spatially constant, and we set it equal to the
accretion rate,
,
4r2 v M
(8.85)
i.e. the mass accepted by the central body per unit time, which is
defined with a negative sign because v < 0; likewise for a stationary, spherically-symmetric flow, Eulers equation says
~ v=v
(~v )~

dv
1 dP GM
=

,
dr
dr
r

(8.86)

because the gravitational potential is


=

GM
r

d
~ = GM ;
= ~er
dr
r2

(8.87)

we now further assume a polytropic equation-of-state, i.e. P


with the adiabatic index 1 5/3; moreover, the pressure
gradient can be written as
dP dP d
d
=
= c2s
;
dr
d dr
dr

(8.88)

we now use these three equations and rewrite them; first, we can
conclude from the continuity equation
"
#
1 d(r2 v)
1 d(r2 v)
2 d
=0 = 2
+r v
r2 dr
r
dr
dr
2
d
d(r v)
+v ,
(8.89)
= 2
r dr
dr
thus

1 d(r2 v)
1 d
= 2
;
dr
r v dr

(8.90)

with this result, Eulers equation can be written as


v

dv c2s d
dv c2 d(r2 v) GM
+
= v 2s
+ 2 =0,
dr dr
dr r v dr
r

(8.91)

where we can use

dv 1 dv2
v =
dr 2 dr
to arrive at the more convenient equation
!
!
c2s dv2
2c2s r
GM
1
1 2
= 2 1
;
2
v dr
r
GM

(8.92)

(8.93)

CHAPTER 8. FLOWS OF IDEAL AND VISCOUS FLUIDS


far away from the central mass, the term
!
2c2s r
1
GM

99

(8.94)

is negative, thus the right-hand side of (8.93) is positive; since


d(v2 )/dr < 0 in order to have inflowing gas, the factor 1 c2s /v2
must be negative, and hence the flow at large radii must be subsonic, v < cs ;
the gas flows towards the accreting object and will reach a point
at sufficiently small rc where
!
2c2s r
1
=0,
(8.95)
GM
from which the critical radius rc can be read off to be
rc =

GM
;
2c2s

(8.96)

inserting the sound speed from


dP d[P0 (/0 ) ] P0 1
=
=
d
d
0
P nkT
kT
=
=
=

c2s =

(8.97)

yields
rc =

GMm
;
2kT

(8.98)

this critical radius is typically far beyond the radius of the central
object, i.e. the accretion flow passes into the supersonic regime
there;
Eulers equation can be integrated; first,
"
! #
Z
Z
Z

dr dP
dr d
P0
d
P0
=
=
dr 2
dr
dr
0
dr

! 0
Z
1
1
P0
d
P0
=
dr
=
, (8.99)

dr 1
0
( 1)0
then, we can use the squared sound speed
P0 1

(8.100)

c2s
dr dP
=
,
dr
1

(8.101)

c2s =
from (8.97) to find
Z

where c2s is a function of radius r, of course;

CHAPTER 8. FLOWS OF IDEAL AND VISCOUS FLUIDS

100

thus, Eulers equation implies


c2
v2
GM
+ s
= const. C ;
2 1
r

(8.102)

since v 0 at r , the constant must equal


c2s ()
C=
,
1

(8.103)

but on the other hand we must satisfy at the critical radius


v = cs

at r = rc =

GM
,
2c2s

(8.104)

so that we can conclude from (8.102)


c2s () c2s (rc ) c2s (rc )
=
+
2c2s (rc ) ,
1
2
1
or
c2s

(8.105)

!
c2 ()
1
1
= s
2+
,
1
2
1

(8.106)

and thus
"

1
cs (rc ) = cs ()
1 3/2( 1)

#1/2

2
;
5 3

= cs

(8.107)

because of the proportionality c2s 1 , the density follows from


"
#2/(1)
cs (rc )
(rc ) = ()
;
(8.108)
cs ()
this also specifies the accretion rate, since
= 4r2 v = const. = 4rc2 (rc )cs (rc ) ,
M

(8.109)

and therefore
=
M

4rc2 ()

2
5 3

!1/(1)

2
cs ()
5 3

()
2
= G M 3
cs () 5 3

!1/(1)+1/22

2
()
= G M 3
cs () 5 3

!(53)/[2(1)]

!1/2

(8.110)

the solution for the velocity finally follows from mass conservation,

4r2 v = M
v =

4r2 (r)

(8.111)

M
cs ()
4r2 () cs (r)
"

#2/(1)
;

CHAPTER 8. FLOWS OF IDEAL AND VISCOUS FLUIDS

101

inserting this into equation (8.102) yields the solution for v(r),
and mass conservation then yields (r); indeed, the maximum
accretion rate at given r is reached exactly when that radius is
critical, r = rc ; i.e. if the flow velocity reaches the sound velocity
there;

Chapter 9
Shock Waves and the Sedov
Solution
9.1
9.1.1

Steepening of Sound Waves


Formation of a Discontinuity

as long as perturbations remain small in the sense that  ,


P  P and v  v, they propagate with the sound speed
and retain their initial (arbitrary) shape, as guaranteed by the
dAlembert equation governing them;
in a polytropic gas, the pressure is P and thus the sound
speed cs (1)/2 increases if the density of the medium increases; waves thus run faster in denser regions and can even pass
less dense parts of the waves;
the density varies along the wave; in denser regions, the wave
propagates faster, which makes the wave front steepen; however,
since the density must be unique at any location, this behaviour
must lead to discontinuities, i.e. to sharp density jumps;
we analyse this behaviour in the simple case of an isentropic flow,
which means s = const. and the same everywhere in the fluid; in
this case, the pressure is a function of density alone, P = P(),
and likewise the velocity along the wave can be written as a function of density alone, v = v(); for a wave propagating into the
positive x direction, the continuity equation requires
(v)
d(v)
+
=0=
+
;
(9.1)
t
x
t
d x
similarly, Eulers equation requires
!
v
v 1 P
v
1 dP v
+v +
=0=
+ v+
;
t
x x
t
dv x
102

(9.2)

further reading: Shu, The


Physics of Astrophysics, Vol II:
Gas Dynamics, chapter 15 and
17; Padmanabhan, Theoretical
Astrophysics, Vol. I: Astrophysical Processes, sections 8.10
8.12; Landau, Lifshitz, Theoretical Physics, Vol VI: Hydrodynamics, chapter IX

CHAPTER 9. SHOCK WAVES AND THE SEDOV SOLUTION 103


the total change in density is
d =

dt + dx ,
t
x

(9.3)

i.e. the change in x at constant is


x
t

=
t

!1
=

d(v)
dv
=v+ ,
d
d

(9.4)

where (9.1) was used;


the change of x at constant v follows similarly,
v
v
dv = 0 = dt + dx
t
x
!
!
!1
x
v v
=
,
t v
t x

(9.5)

or, with the help of Eulers equation (9.2),


!
x
1 dP
=v+
;
t v
dv

(9.6)

however, since is a function of v alone, the change of x with constant must equal the change of x with constant v, and therefore
the two partial derivatives (9.4) and (9.6) must equal,
v+

1 dP
1 dP d
dv
=v+
=v+
;
d
dv
d dv

(9.7)

from this, we conclude


dv
d

!2
=

c2s
,
2

(9.8)

hence the velocity can be expressed as


Z
Z
cs d
dP
v=
=
;

cs
for our polytropic equation of state,
!

P = P0
and
0

P=

kT
,
m

(9.9)

(9.10)

and the squared sound speed is


c2s

P
1

=
= P0 = P0

0
0

!
=

kT
,
m

(9.11)

CHAPTER 9. SHOCK WAVES AND THE SEDOV SOLUTION 104


thus c2s T , and according to (9.10) the temperature scales with
the density as
T 1

or T 1/(1) ,

(9.12)

which can be used to express the density in terms of the sound


speed,
!2/(1)
cs
= 0
;
(9.13)
c0
the velocity follows from (9.6), which can be combined with (9.8)
to find x,
!
x
1 dP
1 dP d
=v+
=v+
= v cs ,
(9.14)
t v
dv
d dv
which yields
x(v, t) = (v cs )t + f (v) ,

(9.15)

where f (v) is an arbitrary function of v to be specified by the


boundary conditions;
with
!2/(1)
!

cs
d

= d ln
= d ln

0
c0
2 dcs
cs
2
,
d ln =
=
1
c0 1 cs

(9.16)

the sound speed can be rewritten from (9.9)


Z
cs d
2(cs c0 )
1
v=
=
cs = c0
v ; (9.17)

1
2
putting this into (9.15) finally yields
!
+1
x = c0 +
v t + f (v) ;
2

9.1.2

(9.18)

Specific Example

to give an example, we study an infinitely long pipe along the


x direction which is closed from the left side with a piston and
filled with gas from the right side; until t = 0, the piston be at
rest at x = 0, and it be accelerated into the pipe with a constant
acceleration a afterwards;

CHAPTER 9. SHOCK WAVES AND THE SEDOV SOLUTION 105


the velocity and position of the piston are thus
vp = at ,

a
xp = t2 ,
2

(9.19)

and since the velocity of the gas has to equal the velocity of the
piston vp at the pistons position xp , we find from (9.18) at the
position of the piston
!
a2
+1
t = c0 +
at t + f (v) ;
(9.20)
2
2
using t = v/a, we can solve for f (v),

v  v
c0 v v2
f (v) = c0 +
=

,
2 a
a
2a

(9.21)

which gives us the general solution for the relation between x and
t to the right of the piston
!
c0 v v2
+1
v t

;
(9.22)
x = c0 +
2
a
2a
the velocity at position x and time t is thus determined by
!
v2
c0 1
+

t v (c0 t x) = 0 ,
(9.23)
2a
a
2
which can be solved for v to yield
s

!
!2

1 +1
+1
v =
at c0
at c0 + 2a(c0 t x) ;

2
2
(9.24)
this is the velocity of the gas for all points x to the right of the
piston, i.e. for x at2 /2;
for x = 0 and t = 0, the velocity must vanish, which selects from
the two branches of the solution (9.24) the one with the positive
sign,
s

!
!2

1 + 1
+1
v =
at c0
at c0 + 2a(c0 t x) ;

2
2
(9.25)
a discontinuity is formed where the velocity changes suddenly,
which is where
!
!
v
x
=
=0
(9.26)
x t
v t

CHAPTER 9. SHOCK WAVES AND THE SEDOV SOLUTION 106


holds; physically, this means that parts of the wave with different
velocities meet at the same position; differentiating (9.22) with
respect to v, we see that this happens in our example when
+1
c0 v
t
=0
(9.27)
2
a
a
at the time
2
tc =
(c0 + v) ;
(9.28)
a( + 1)
there, the velocity drops to zero, v = 0, and thus
2c0
;
(9.29)
tc =
a( + 1)
specifically, setting = 5/3 and a = c0 / with a characteristic
acceleration time scale , the gas velocity is
s

!
!2
!

3c0 4t
4t
10 t
x
v=
1 +
1 +

, (9.30)
5 3
3
3 c0
where x must obey
x

at2 c0 t2 c0  t 2
=
=
2
2
2

1  t 2
x

;
c0 2

(9.31)

now, we define
x
2
t
v

;
; ,
c0 2

c0
which enables us to write (9.22) as
s

!
!2
!

4
10
2
3 4
=
1 +
1 +
,
5 3
3
3
2

(9.32)

(9.33)

where 0 is now the distance from the piston in units of c0 ;


the discontinuity occurs at the time
3
3
tc =
or c = ;
4
4
since the velocity must vanish there,
v=0=,

(9.34)

(9.35)

the position of the discontinuity is determined by


c

c2
= c
2

15
;
32

(9.36)


1

v = c0 1 +
2
3

(9.37)

c =

the sound speed is


cs = c0 +
and thus the density is

3
= 1+
;
0
3

(9.38)

CHAPTER 9. SHOCK WAVES AND THE SEDOV SOLUTION 107

9.2
9.2.1

Shock Waves
The Shock Jump Conditions

shock waves occur when a supersonic flow hits an obstacle, for


instance because it impinges on a solid, resting body, or because
a body moves at supersonic speeds through the fluid; in the immediate vicinity of the body, the velocity must drop tp zero, and
because the flow is supersonic, no information on the obstacle can
propagate upstream against the flow;
we approximate the discontinuity as a plane and consider a supersonic fluid flow perpendicularly hitting the plane from the left; let
1 , P1 and v1 be the density, pressure and velocity of the fluid to
the left of the discontinuity, and 2 , P2 and v2 be to its right;
deriving Eulers equations, we had identified the following currents:
~
mass current
 v2

~v
+ w ~v energy current
(9.39)
2
i j
ij
v v + P momentum current
when specialised to the situation considered here, in which the
discontinuity is perpendicular to the flow along the x axis, the
conservation of these currents requires the following conditions
1 v1 = 2 v2
!
~v22
+ w1 1~v1 =
+ w2 2~v2
2
2
1 v21 + P1 = 2 v22 + P2 ;

~v21

(9.40)

the enthalpy and the sound speed for a polytropic gas are
w=

P
,
1

c2s =

P
;

(9.41)

writing the velocity left of the discontinuity as


v1 = M1 cs

(9.42)

with the Mach number M1 > 1, the equations (9.40) can be rewritten in the following way
( + 1)M21
2
v1
=
=
2
1
( + 1) + ( 1)(M1 1) v2
( + 1) + 2(M21 1)
P2
=
,
P1
+1

(9.43)

CHAPTER 9. SHOCK WAVES AND THE SEDOV SOLUTION 108


and for an ideal gas with P = kT/m, the temperature ratio is
T 2 P2 1
=
;
T 1 P1 2

(9.44)

by construction, the Mach number M1 > 1, which implies 2 >


1 because
( + 1) + ( 1)(M21 1) = 2 + ( + 1)M21 2M21
= 2(1 M21 ) + ( + 1)M21
(9.45)
< ( + 1)M21 ,
and therefore the density ratio from (9.43) is larger than unity;
correspondingly, v2 < v1 , P2 > P1 and T 2 > T 1 ;
in the limiting case of a highly supersonic flow, M1 and
2 + 1
=
;
M1 1
1
lim

(9.46)

for a gas with adiabatic index = 5/3, the maximum density ratio
is therefore
+1
=4,
(9.47)
1
which is called the maximum shock strength; in the same limit,
P2
,
P1

9.2.2

T2
;
T1

(9.48)

Propagation of a One-Dimensional Shock Front

now we consider a fluid pipe with a piston, which remains at rest


at x = 0 until t = 0 and is then instantly accelerated to a velocity
u into the positive x direction;
a discontinuity forms at t = 0 which propagates with a velocity vs
to the right; then, there exists a region ahead of the shock where
the density, pressure and temperature still have their original values 1 , P1 and T 1 ; in the region between the shock and the piston,
the gas moves with the velocity u of the piston; the difference of
the velocities between the two regions is thus u;
in order to use the jump conditions derived before, we need to
transform into a coordinate frame in which the shock is at rest; in
this (primed) frame, the gas velocities are obviously
v01 = vs ;

v02 = u vs = u + v01 ,

(9.49)

and the velocity difference must remain the same, of course,


v02 v01 = u

(9.50)

CHAPTER 9. SHOCK WAVES AND THE SEDOV SOLUTION 109


eliminating the Mach number M1 , the jump conditions (9.43) for
the density and the pressure P can be combined to read
2 P2 ( + 1) + P1 ( 1)
=
;
1 P2 ( 1) + P1 ( + 1)

(9.51)

the matter current is j = 1 v01 , thus


2 02
j2 = 21 v02
1 = 2 v2 ,

(9.52)

and therefore
 1h
i
1  2 02
2
02
1 v1 + 22 v02
=
j
+
(P

P
)
+

v
, (9.53)
j2 =
2
1
1
1
2
2
2
2
2
where we have used the jump condition for the momentum from
(9.40); we thus obtain
"
#
1 2
2
2 1
j =
j + (P2 P1 )1 + j
(9.54)
2
2
or
j2 =

P1 P2
1 2 ;
1 2

(9.55)

the velocity difference can be written as follows:


v02

v01

1 v02 1 v01
1 1 0
=
2 v2 1 v01
= u=
1
1 2
1 2
= j
;
1 2

with (9.55), this turns into


"
#1/2
(P1 P2 )(1 2 )
u=
1 2
according to (9.51), the density ratio is
!
1 2
1 1
=
1
1 2
1 2
"
#
1 P1 ( + 1) + P2 ( 1)
1
=
1 P1 ( 1) + P2 ( + 1)
2
P1 P2
=
,
1 P1 ( 1) + P2 ( + 1)
which allows us to write the velocity difference as
s
2
u = cs |1 |
,
( 1) + ( + 1)

(9.56)

(9.57)

(9.58)

(9.59)

where P2 /P1 is the pressure ratio, and we have expressed the


ratio P1 /1 in terms of the sound speed cs in the unshocked gas
ahead of the piston from (9.41), P1 /1 = c2s /;

CHAPTER 9. SHOCK WAVES AND THE SEDOV SOLUTION 110


this can be re-written as a quadratic equation for the pressure ratio
,
# "
#
"
( 1)u2
( + 1)u2
2
+ 1
=0
(9.60)
2+
2c2s
2c2s
which has the solutions
( + 1)u2
= 1+
(9.61)
4c2s
s
!2
!
( 1)u2
( + 1)u2

1
;
1+
4c2s
2c2s
the pressure ratio needs to exceed unity, 1, which excludes
the negative branch; the solution for the pressure ratio can thus be
simplified to
s
( + 1)u2 u
( + 1)2 u2
=1+
+
1
+
;
(9.62)
4c2s
cs
16c2s
note that, if the piston is at rest, u = 0 and = 1, as expected;
equations (9.56) and (9.58) together yield v01 ,
u = 2v01

1
,
1 + ( + 1)

which can be simplified by means of (9.59) to read


cs p
v01 = p
1 + ( + 1) ;
2

(9.63)

(9.64)

this is the velocity of the unshocked gas in the rest frame of the
shock front; obviously, the velocity of the shock front in the rest
frame of the unshocked gas is
vs = v01 ,

(9.65)

hence (9.65) also yields the (negative) velocity of the shock front
in our laboratory system;
the physical conditions in the unshocked gas, expressed by
(1 , P1 , T 1 ) and the velocity u of the piston, first yield P2 /P1
from (9.62), from which 2 and T 2 immediately follow using the
shock jump conditions; the velocity of the shock relative to the
velocity of the piston is given by (9.64); if u  cs , we find

( + 1)u2
,
2c2s

(9.66)

and the shock velocity becomes


+1
u&u,
(9.67)
2
i.e. for a gas with = 5/3, the shock moves 33% faster than
the piston;
vs

CHAPTER 9. SHOCK WAVES AND THE SEDOV SOLUTION 111

9.2.3

The Width of a Shock

one of our assumptions was that the energy is conserved across


the discontinuity of a shock front; this is not necessarily so because it can be transported away due to radiative losses, friction,
diffusion of fast particles and similar processes;
if the energy remains conserved, the shock front is called adiabatic; otherwise, viscous and thermal effect need to be taken into
consideration; energy conservation then demands
!
~v2
~
const. = ~v
+ w T
(9.68)
2
~ ~v) + 2 ~v(
~ ~v) 2(~v )~
~ v,
+ ~v(
3
which can be specialised to our case of a one-dimensional shock
front,
!
!
T
4
v
v2
v
+w
v = const. ;
(9.69)
2
x
3
x
we now assume that dominates over and put in coarse approximation = 0; moreover, the viscous friction term must be
comparable to the kinetic energy if it is to play an important role;
thus
v2 4 v
(9.70)
v v ;
2
3 x
if the velocity changes by v across a distance x, we estimate
v

v
v
v
v
v
,
x
x
x

(9.71)

from which we can estimate x,

v2 4 v

2
3 x

8 v
;
3v v

(9.72)

across a strong shock, v v, which yields the shock width


x

8
;
3v

(9.73)

since the viscosity is v, the shock with turns out to be x


8/3, i.e. it is a factor of order unity times the mean-free path
length;

CHAPTER 9. SHOCK WAVES AND THE SEDOV SOLUTION 112

9.3

The Sedov Solution

9.3.1

Dimensional Analysis

one example for a shock wave is given by an explosion, i.e. by an


event in which in very short time energy is being released within a
very small volume; we consider such an event under the following
simplifying assumptions: (1) the shock is very strong, meaning
that the pressure of the surrounding medium can be neglected,
P1  P2 ; (2) the energy E is released instantaneously; and (3)
the energy of the surrounding material is negligible compared to
E, i.e. the explosion energy dominates that of the surroundings;
and finally (4) the gas be polytropic with an adiabatic index ;
under these conditions, our shock jump condition for the density
is
1 P1 ( + 1) + P2 ( 1) 1
=

,
(9.74)
2 P1 ( 1) + P2 ( + 1) + 1
which implies that 1 and 2 are completely determined by one
another; the behaviour of the shock must thus be entirely determined by the explosion energy E and the surrounding matter density 1 ;
if we now consider the shock at a time t when it has reached
the radius R(t), the only quantity with the dimension of a length,
which can be formed from E, t and 1 is
Et2
1

!1/5
;

(9.75)

which makes us set


R(t) = R0

Et2
1

!1/5
(9.76)

with a dimension-less constant R0 which remains to be determined;


the shock velocity is obviously
dR
E
vs =
= R0
dt
1

!1/5

2t2/51 2 R
=
;
5
5t

(9.77)

we now use the jump conditions which we had obtained for the
piston in the tube; first, the velocity of the piston is, according
to (9.67),
2vs
u=
,
(9.78)
+1

CHAPTER 9. SHOCK WAVES AND THE SEDOV SOLUTION 113


from which (9.66) yields for the pressure within (behind) the
shock
!1
21 v2s
2v2s P1
( + 1) u2
=
P2 = P1
=
P
,
(9.79)
1
2
c2s
+ 1 1
+1
where the sound speed (9.41) in the surrounding, unshocked
medium was inserted;
our earlier expression (9.74) for the density shows that the density
inside the shock remains constant, because 1 is constant; since
the shock velocity (9.77) drops with time like
vs t2/51 = t3/5 ,

(9.80)

the pressure inside the shock drops like


P2 t6/5 ,

(9.81)

and the velocity of the gas behind the shock is


u vs t3/5 ;

(9.82)

we can interpret these relations as follows: a shock wave driven


by the release of the energy E, which propagates outward with
the time-dependent radius R(t), collects material with mass
M 1 R3 ;

(9.83)

this material is accelerated from zero to a velocity R/t, such


that the kinetic energy
R5
1 2
(9.84)
t
is put into the collected material; the energy of the material behind
(within) the shock is thus approximately
1 R 2 R3 1

R5
;
t2

equating this to the energy E, we immediately find


!1/5
Et2
R=
,
1

(9.85)

(9.86)

i.e. the scaling relation (9.76) simply expresses energy conservation within the shock;
we now know how the velocity, the radius, the pressure and the
density at the shock; they are completely determined by the release of an amount of energy E into surrounding material with
the density 1 whose energy can be neglected;

CHAPTER 9. SHOCK WAVES AND THE SEDOV SOLUTION 114

9.3.2

Similarity Solution

the external radius of such an explosion is given by (9.86), which


suggestes introducing
r

(9.87)
R
as a dimension-less radial variable; we will now use to express the radius in v(r, t), (r, t) and P(r, t) and solve the hydrodynamic equations to determine the properties of the gas everywhere within the shock;
the velocity at the shock is given by (9.76); imitating this behaviour, we put
2r
v(, t) =
V()
(9.88)
5t
with a dimension-less function V() which needs to be determined; the gas velocity at the inner rim of the shock, given by
(9.78), requires that V() satisfy the boundary condition
V(1) =

2
;
+1

(9.89)

similarly, we use the ansatz


= 1G()

(9.90)

for the density and must, because of (9.74), satisfy the boundary
condition
+1
G(1) =
(9.91)
1
for the as yet unknown function G();
we finally express the pressure by the sound speed, using (9.42)
together with (9.74), (9.77) and (9.79) to write
c2s

21 2 2( 1) 2 R
P2
=
=
v =
2
( + 1)2 s
( + 1)2 5 t

!2
,

(9.92)

which justifies the ansatz


c2s =

4 r2
Z() ,
25 t2

(9.93)

where Z() must satisfy the boundary condition


Z(1) =

2( 1)
;
( + 1)2

(9.94)

CHAPTER 9. SHOCK WAVES AND THE SEDOV SOLUTION 115


now, we can use energy conservation again to relate Z and V; the
energy E within the shock must remain constant because no energy can flow at supersonic velocities; since the solutions, however, have to be self-similar because they are expressed by the
dimension-less radius , energy must also be conserved within
any other sphere with , 1, because , 1 gives the shock position at another than the considered instant of time; consequently,
we can set up the energy balance for any sphere and require that
the energy remain constant;
per time interval dt, a sphere with radius r loses the energy
!
v2
2
4r v
+ w dt ,
(9.95)
2
while it gains energy by growing by an amount
!
2
2 2r
dt ,
4r vs dt = 4r
5t
incorporating the additional amount of energy
!
!
v2
2 2r
4r
dt  +
;
5t
2
the energy balance then implies
!
!
v2
2r v2
v
+w =
+ ,
2
5t 2

(9.96)

(9.97)

(9.98)

and since the enthalpy is


w=

c2
P
+P
=
= s ,

1 1

(9.99)

the thermal energy per unit mass is


=

c2s
1 P

=
=
;
1 ( 1)

this implies with (9.98)


!
!
c2s
c2s
2r v2
v2
v
+
=
+
,
2 1
5t 2 ( 1)

(9.100)

(9.101)

where we now insert the ansatze for v and c2s , (9.88) and (9.93);
the result can be written as
!
Z
V2
Z
V2
V
+
=
+
,
(9.102)
1
2
( 1)
2
from which follows
Z=

( 1)(1 V)V 2
;
2(V 1)

(9.103)

CHAPTER 9. SHOCK WAVES AND THE SEDOV SOLUTION 116


the hydrodynamic equations which we now have to solve are of
course the continuity, Euler, and energy conservation equations;
when specialised to our radially symmetric problem, they read
(v) 2v
+
+
= 0
t
r
r
v
v
1 P
+v
=
t
r
r
!
!
P

ln = 0 ,
+v
t
r

(9.104)

where the conservation of entropy was used instead of the energy


conservation equation;
inserting G and V into the continuity equation and using
d
2 d

=
=
t t d
5t d

(9.105)

yields after some straightforward manipulation


V 0 (1 V)

G0
+ 3V = 0
G

(9.106)

noticing that
1
P
P 1 c2s 1 4r2 1
= = =
ZG1

25t2

(9.107)

and dropping irrelevant constants, the logarithm in (9.104) can be


re-written as
!
P
r2
1
ln = ln 2 ZG
,
(9.108)

t
and substituting this into the entropy-conservation equation
(9.104) yields
Z 0
G0 5 2V
+ (1 )
+
=0;
Z
G
1V

(9.109)

eliminating G0 /G from the continuity equation (9.106), this latter equation becomes
 5 2V
Z 0 1
+
3V + V 0 +
=0,
Z
1V
1V
which is supplemented by (9.103), which implies
!
Z 0
2
1

V 0 ;
Z
V 1 V V 1

(9.110)

(9.111)

CHAPTER 9. SHOCK WAVES AND THE SEDOV SOLUTION 117


taken together, (9.110) and (9.111) yield the single equation for
V,
V 0 (1 3)V 2 + (8 1)V 5
=
;
(9.112)
V
( + 1)V 2 2( + 1)V + 2
this ordinary first-order differential equation can directly be integrated in closed form after separating variables, using the boundary condition (9.89); having found V(), Z follows from (9.103),
and G from (9.107);

Chapter 10
Instabilities, Convection, Heat
Conduction, Turbulence
10.1

Rayleigh-Taylor Instability

instabilities occur in many different forms in hydrodynamical


systems; their analysis always proceeds according to the same
scheme: one starts from an equilibrium configuration, perturbs
it slightly, i.e. in linear approximation, decomposes the perturbations into plane waves, exp[i(t ~k ~x)] and derives the dispersion relation (k); imaginary frequencies signal the onset of
instabilities;
we first study the situation in which two fluids with the densities
1 and 2 are separated by a plane; we choose that plane to be the
x-y plane and assume that gravity is directed into the negative z direction with the acceleration g, corresponding to the gravitational
potential = gz; finally, the perturbation of the separating plane
be described by a function (x), i.e. perturbations are assumed to
be independent of y, without loss of generality;
since ~v is assumed to be small, we can neglect the curl term in
Eulers equation (8.6) and assume that the velocity is the gradient
of a velocity potential ; then, we can use Bernoullis law in the
form (8.41), where we ignore the term ~v2 /2 because it is of second
order in ~v; thus,

+ w + gz = 0 ;
(10.1)
t
for an incompressible fluid, and the enthalpy turns into
w=

P
P
;
1

118

(10.2)

further reading: Shu, The


Physics of Astrophysics, Vol II:
Gas Dynamics, chapter 8; Padmanabhan, Theoretical Astrophysics, Vol. I: Astrophysical
Processes, sections 8.138.15;
Landau, Lifshitz, Theoretical
Physics, Vol VI: Hydrodynamics, chapter III

CHAPTER 10. INSTABILITIES, CONVECTION, HEAT CONDUCTION, TURBULENCE119


multiplying (10.1) with thus yields

+ P + gz = 0
t

P = gz

;
t

pressure equilibrium at the boundary then implies






= 2 g + 2
,
1 g + 1
t 1
t 2

(10.3)

(10.4)

which constrains the boundary by



!
1


1
;
2
=
g(1 2 )
t 2
t 1

(10.5)

since the velocity must be continuous across the boundary, the


velocity components in z direction must also agree on both sides
of the boundary, thus




=

at z = ;
(10.6)
z 2
z 1
finally, the velocity components in z direction can to first order be
identified with the displacement of the boundary,

=
;
t
z

(10.7)

differentiating the equation (10.5) for the boundary with respect


to t, this yields


2
2

= 2 2 1 2
g(1 2 )
(10.8)
z
t 2
t 1
~ ~v = 0, implies that the velocity potential
the incompressibility,
~ 2 = 0, and since is indepen satisfies Laplaces equation,
dent of y by construction, we can set
= f (z) cos(t kx) ;

(10.9)

the Laplace equation then demands


d2 f
cos(t kx) f k2 cos(t kx) = 0 ,
dz2

(10.10)

thus f satisfies an oscillator equation with the usual exponential


solutions,
d2 f
k2 f = 0 f ekz ;
(10.11)
dz2

CHAPTER 10. INSTABILITIES, CONVECTION, HEAT CONDUCTION, TURBULENCE120


let now h1 and h2 be the heights of the layers, then the velocity
needs to vanish at both z = h1 and z = h2 ; this specifies the
solution
1 = A1 cosh [k(z + h1 )] cos(kx t) ,
2 = A2 cosh [k(z h2 )] cos(kx t) ;

(10.12)

if we insert these solutions into the equation (10.8) for the pressure balance, we find
g(1 2 )A1 sinh[k(z + h1 )] k cos(kx t) =
1 A1 cosh[k(z + h1 )]2 cos(kx t)
(10.13)
+ 2 A2 cosh[k(z h2 )]2 cos(kx t) ;
from (10.13), we obtain the ratio A1 /A2 ,
A2 g(1 2 ) k sinh[k(z + h1 )] + 1 2 cosh[k(z + h1 )]
=
,
A1
2 2 cosh[k(z h2 )]
(10.14)
and a similar expression follows if 1 and 2 are swapped; equating both yields the dispersion relation
2 =

kg(1 2 )
;
1 coth kh1 + 2 coth kh2

(10.15)

if 2 > 0 as required for a stable situation, 1 > 2 is obviously


necessary, which means that the specifically heavier fluid must
lie below the specifically lighter fluid; in this case, perturbations
propagate as waves along the boundary between the fluids;
if the density of the upper fluid is small compared to the lower,
we can approximate 2 = 0 and have
2 =

kg1
= kg tanh kh1 ,
1 coth kh1

(10.16)

which gives the frequency of waves on a fluid under the influence


of gravity whose surface is being perturbed; in the limiting case
of low depth, h1  k1 , this simplifies to
2 k2 gh1 ,

(10.17)

while for large depth, h1  k1 and


2 kg ;

(10.18)

these are the limiting cases of waves on shallow or deep water;

CHAPTER 10. INSTABILITIES, CONVECTION, HEAT CONDUCTION, TURBULENCE121


if 2 > 0 and kh1  1 as well as kh2  1, we find
2 kg

1 2
;
1 + 2

(10.19)

this is the limiting case of very deep layers; in the limit of long
waves, kh1  1 and kh2  1, the dispersion relation becomes
2 k2

g(1 2 )h1 h2
;
1 h2 + 2 h1

(10.20)

this linear stability analysis shows that becomes imaginary if


the specifically heavier fluid lies on top of the specifically lighter
one, 1 < 2 , which is intuitively obvious;

10.2

Kelvin-Helmholtz Instability

we now consider a situation in which one fluid flows with a velocity ~v parallel to the surface of a fluid at rest, for instance like
wind over a lake;
we choose the coordinate system such that the z axis is orthogonal
to the boundary surface and direction of motion (i.e. the direction
of ~v) points into the x direction;
writing Eulers equation in the form
~
d~v
P
=
dt

(10.21)

and taking the divergence, we see that the pressure must satisfy
Laplaces equation
~ 2P = 0

(10.22)
~ ~v = 0 and = const.;
if the fluid is incompressible,
as before, the ansatz P = f (z) exp[i(kx t)] leads to the oscillator equation
d2 f (z)
k2 f (z) = 0 ,
(10.23)
dz2
with the solutions f (z) exp(kz); the exponentially growing
solution exp(kz) is ruled out physically because it diverges at
large distances from the boundary surface, and thus the pressure
perturbation must behave as
P2 ekz ei(kxt)
above the boundary surface;

(10.24)

CHAPTER 10. INSTABILITIES, CONVECTION, HEAT CONDUCTION, TURBULENCE122


to first order in the velocity perturbation vz along the z axis, Eulers equation reads
vz
vz k P2
+v
=
,
t
x
2

(10.25)

which, with vz exp[i(kx t)], implies


ivz + ikvvz =
and thus
vz =

kP2
,
2

kP2
;
i2 (kv )

(10.26)

(10.27)

let now again (x, t) be the boundary surface between the two
fluids, then we must have
d

= vz =
+v
(10.28)
dt
t
x
to linear order, and the ansatz exp[i(kx t)] implies
i + ikv = vz = i(kv ) ,

(10.29)

thus the pressure perturbation P2 from (10.27) can be written as


P2 =

2
(kv )2 ;
k

(10.30)

on the other side of the boundary surface, we have v = 0 and must


choose the solution f (z) exp(kz); inserted into Eulers equation,
this yields
vz
k P1
=
t
1

vz =

k P1 ik P1
=
;
i1
1

(10.31)

the boundary surface then satisfies the equation

or

= i = vz ,
t

(10.32)

1 2
P1 =
;
k

(10.33)

there must be pressure balance at the boundary, P1 = P2 , therefore


1 2 + 2 (kv )2 = 0 = (1 + 2 )2 22 k + 2 k2 2 , (10.34)
and this equation has the solutions
q
22 kv 422 k2 v2 4(1 + 2 )2 k2 v2
=
2(1 + 2 )

kv

=
2 i 1 2 ;
1 + 2

(10.35)

CHAPTER 10. INSTABILITIES, CONVECTION, HEAT CONDUCTION, TURBULENCE123


for 1 , 0 , 2 , always has an imaginary part, i.e. the perturbation grows; this so-called Kelvin-Helmholtz instability can be
damped by buoyancy forces, if 1 > 2 ; of course, a gravitational
field must then be taken into consideration;
the time scale for the perturbation to grow is obviously given by


=()

1

1 + 2
;

kv 1 2

(10.36)

2
1

(10.37)

if 1  2 ,
ikv

follows, which holds for example for wind blowing over water;

10.3

Thermal Instability

we assume that the system under consideration gains energy by


heating and loses energy by cooling; the net cooling rate, i.e. the
net energy loss per unit time, be
L(, T ) ;

(10.38)

if the system is in thermal equilibrium, we require L(, T ) = 0,


which implicitly defines a relation between and T ; for thermal
bremsstrahlung, for example,

L(, T ) = C T (heating) ;
(10.39)
this cooling function L(, T ) can adopt various forms, in particular because cooling processes are often related to thermal occupation numbers and atomic or molecular excitations; because of the
Boltzmann factor, sometimes small temperature changes can give
rise to large changes in occupation numbers, and the atomic or
molecular energy levels introduce discrete thresholds; the curve
L(, T ) = 0 thus typically looks as shown in the figure (to be
inserted);
at the same time, let the system be in pressure equilibrium with
its surroundings, i.e. the pressure P be externally regulated; for
an ideal gas, we have
P T ,
(10.40)
such that pressure equilibrium may be represented by a straight
inclined line; in this example, the curve P = const. intersects the
curve L = 0 in three points where both mechanical and thermal
equilibrium are possible;

CHAPTER 10. INSTABILITIES, CONVECTION, HEAT CONDUCTION, TURBULENCE124


if the system moves along the curve P = const., it gets out of thermal equilibrium; the idea behind that is that mechanical is usually
established much faster than thermal equilibrium such that a system more likely stays in mechanical than in thermal equilibrium if
and T are changed, and thus it remains on the curve P = const.;
if, at the point A, the density is increased, the system moves
towards the lower right along the dashed line and thus into the
regime where L > 0; here, the energy losses are larger than the
energy gain, the temperature decreases further, the density grows
further, and the system moves further away from thermal equilibrium, it is thermally unstable;
if the same happens in point B, the cooling function becomes negative, L < 0, the system heats up and moves back into the equilibrium point;
we now consider a simple model for this thermal instability; the
fundamental equations are
~
+ (~v) = 0
t

~ 2
~
~v v
~ ~v) = P
+
~v (
t
2

#
"
s
~
+ (~v )s
= L(, T ) ,
T
t

(10.41)

where the specific entropy s is


P
s = cv ln

!
(10.42)

up to a constant which is irrelevant here; L is the net energy loss


per unit mass, and the equation of state is
P=

kT
= RT
m

(10.43)

with the mol number and the gas constant R; the specific heat
capacities at constant volume and constant pressure are
cv =

R
1

(10.44)

and
!
1
R
cp = cv + R = R
+1 =
= cv
1
1

(10.45)

CHAPTER 10. INSTABILITIES, CONVECTION, HEAT CONDUCTION, TURBULENCE125


the equilibrium state has = 0 , T = T 0 , L(0 , T 0 ) = 0 and ~v = 0;
we perturb this state by small deviations , T , ~v and linearise
in these perturbations; this first yields
~
+ (0 ~v) = 0 ,
t

~
~v
P
=
t
0

(10.46)

and, by substituting into the partial time derivative of the first


equation the divergence of 0 times the second equation,
!
2 ~
~v
2 ~ 2
+ 0
P = 0
(10.47)
=
t2
t
t2
we first allow perturbations with P , 0 and ask later for the
conditions for instability under constant pressure;
up to the irrelevant constant, the entropy is
!
P
s = cv ln = cv (ln P ln )



= cv ln(P0 + P) ln(0 + )
( "
!#
"
!#)
P

= cv ln P0 1 +
ln 0 1 +
P0
0
#
"

= cv ln P0 ln 0 +
P0
0
!
P

= s0 + cv
cp ; (10.48)
= s0 + cv
P0
0
P0
0
this allows us to write the left-hand side of the entropy equation
(10.41) as
!
!

P
~
(T 0 + T )
+ ~v s0 + cv
cp
(10.49)
t
P0
0
!
!

s0
~
= T0
+ ~v s0 + cv
cp
+ T
;
t
P0
0
t
equilibriums requires that
!

~ s0 = 0
T0
+ ~v
t

T0

s0
=0
t

(10.50)

because of ~v0 = 0; this reduces the left-hand side of (10.41) to


!

T0
cv
cp
;
(10.51)
t
P0
0
on the right-hand side, we have
L
L
L(0 + , T 0 + T ) = L(0 , T 0 )
T ; (10.52)
| {z }
T
=0

CHAPTER 10. INSTABILITIES, CONVECTION, HEAT CONDUCTION, TURBULENCE126


since and T are furthermore connected through the equation-ofstate, we can write
!
!
L
L
d +
dT
dL =
T
T
! "
!
!
#
!
L

L
=
dP +
dT +
dT
T P T
T P
T
!
!
L
L
=
(10.53)
dT P +
dT ;
T P
T
at constant pressure or constant density ,
dT P = T

d
,

dT = T

dP
,
P

0
T

(10.54)

and thus
cv P cv
L

=
P0 t
0 t
T

!
P

P
;
P0

(10.55)

if we apply the Laplacian to this equation and use (10.47) for


~ 2 P, we find

" 2
#

P0 ~ 2
cv 2 cp =
t
t
0
!
!
L P0 ~ 2
L 2

;
(10.56)
T 0
T P t2
dividing by cv , recalling that
cp
=,
cv

P0
= c2s ,
0

and introducing the abbreviations


!
!
1 L
1 L
, Nv
,
Np
cp T P
cv T

(10.57)

(10.58)

we find the relation


!
2
2
2~ 2
~ 2 Nv ;
cs = N p c2s
t t
t

(10.59)

again, we expand the perturbations into plane waves,


exp[i(kx t)], and thus transform (10.59) to
 i 

h 2 2
cs k 2 = Nv 2 N p c2s k2
t

(10.60)

which yields
(c2s k2 2 )i = Nv 2 N p c2s k2 ;

(10.61)

CHAPTER 10. INSTABILITIES, CONVECTION, HEAT CONDUCTION, TURBULENCE127


this cubic dispersion relation is difficult to solve; in the limiting
case of small wave lengths, c2s k2  2 , we can approximate
i N p

iN p ,

(10.62)

which indicates an unstable solution if N p < 0, since then


eN p t

(10.63)

grows exponentially; thermal instability thus sets in when


!
L
<0;
(10.64)
T P
in the opposite limiting case of very large wave length,
i Nv ,

(10.65)

i.e. instability then requires that


!
L
<0;
T

(10.66)

10.4

Heat Conduction and Convection

10.4.1

Heat conduction

a system can be in mechanical equilibrium, but out of thermal


equilibrium; the most straightforward example is a star which
is being kept in mechanical equilibrium by the balance between
gravity and the pressure gradient, but nonetheless continuously
radiates energy; the entropy equation reads in this case
T

i
ds ~
~ ) + i j v ;
= (T
dt
x j

(10.67)

if the velocity gradient is too small to drive currents, the second


term on the right-hand side can be neglected; from
!
dq
(10.68)
dq = T ds and cp =
dT P
follows
cp dT = T ds

ds = cp d ln T ;

(10.69)

with that, the entropy equation (10.67) can be reduced to


cp

dT
~ 2T
=
dt

or

dT
~ 2T
=
dt

with

; (10.70)
cp

CHAPTER 10. INSTABILITIES, CONVECTION, HEAT CONDUCTION, TURBULENCE128


in analogy to radiative energy transport, we define a conductive
opacity cond through the conductive energy current
F~cond =

c ~
~ ,
(aT 4 ) T
3cond

(10.71)

from which we obtain the relation between heat conductivity


and conductive opacity cond ,
=

4caT 3
;
3cond

(10.72)

if both radiative and conductive energy transport is present, the


effective opacity is
1
1
1
=
+
eff rad cond

eff =

rad cond
;
rad + cond

(10.73)

when we discussed the heat conductivity, we saw that it can be


written as
nv
=
cv ,
(10.74)
3
where is the mead free path of the fluid particles; if we consider
electrons whose mean free path is determined by ions,
=

1
,
ni

(10.75)

where ni and are the number density and the scattering cross
section of the ions;
typically, an electron will approach an ion up to a distance ri
where the kinetic and potential energies equal,
mv2 Ze2

2
ri

ri

2Ze2
;
mv2

(10.76)

the cross section is then


ri2 ,

(10.77)

and we obtain the expression


m2e v4e
m2e
1
ne ve
=
cv
=
3
4ni Z 2 e4 3 4Z 2 e4

!
ne
cv v5e
ni

(10.78)

for the heat conductivity of electrons scattered by ions;


in a thermal electron gas, cv = 3k/2 and
ve =

3kT e
,
me

(10.79)

CHAPTER 10. INSTABILITIES, CONVECTION, HEAT CONDUCTION, TURBULENCE129


which yields the heat conductivity
m2e
k
=
2 4Z 2 e4

ne
ni

3kT e
me

!5/2
(10.80)

for classical (non-degenerate) electrons; if we identify the Thomson cross section T , we can alternatively write

!
!5/2
3 3kc ne kT e
= 2
(10.81)
Z T ni me
with the obvious unit
erg
;
cm s K

[] =

(10.82)

numerically,
5.5 10 Z
12

10.4.2

ne
ni

kT e
1 keV

!5/2
;

(10.83)

Convection

if the temperature gradient is too large, convection sets in: then,


warm, rising bubbles cannot any more cool and return to their
original locations, but continue to rise; we consider the situation,
in which a volume V(P, s) characterised by the pressure P and the
specific entropy s rises against the gravitational force;
we ignore again thermal compared to mechanical adaptation processes because they are typically slower and assume that the bubble with volume V(P, s) rises by an amount z, where its volume
is V(P0 , s); there, its buoyancy force is determined by the volume V(P0 , s0 ) which the bubble would adopt if it had the specific
entropy of its new environment; the situation is stable if the actual volume V(P0 , s) is smaller than the volume V(P0 , s0 ), because
then gravity will dominate the buoyancy force, and the bubble
will then sink down again; we thus have the condition
V(P0 , s0 ) > V(P0 , s)

(10.84)


ds
s = s + z
dz z

(10.85)

for stability;
with

and because of
cp dT = T ds

V
s

!
P

T V
=
cp T

!
>0,
P

(10.86)

CHAPTER 10. INSTABILITIES, CONVECTION, HEAT CONDUCTION, TURBULENCE130


we find
V
V(P , s ) = V(P , s) +
s
0

V
s = V(P , s) +
s
P


ds
z ;
dz z
(10.87)

thus, the stability condition is satisfied if



ds
>0
dz z
in an ideal gas with adiabatic index , the entropy is
h
i
s = cv ln PT /(1)
up to an irrelevant constant; thus
"
#
d ln P
d ln T
ds s dP s dT
;
=
+
= cv
+
dz P dz T dz
dz
1 dz

(10.88)

(10.89)

(10.90)

the condition (10.88) then shows that the temperature gradient


must satisfy
d ln T
1 d ln P

<
(10.91)
d ln z
d ln z
for the gas stratification to be stable against convection; the quantity
1
ad
(10.92)

is often called the adiabatic temperature gradient (nabla adiabatic); using this, we stability condition is written

d ln T
< ad ;
d ln P

(10.93)

when convection sets in, it is a very efficient means of transporting


heat; viscosity hinders the convective energy transport;

10.5

Turbulence

hydrodynamical flows with large Reynolds numbers turn out to be


highly unstable; for high viscosity (low Reynolds number), stable
solutions of the Navier-Stokes equation exist which develop instabilities above a critical Reynolds number
R=

uL
& Rcr ;

(10.94)

a full analysis of such instabilities is very difficult and in general


an unsoved problem; turbulence sets in, in the course of which
energy is being transported from large to small scales until it is
dissipated by the production of viscous heat on sufficiently small
scales;

CHAPTER 10. INSTABILITIES, CONVECTION, HEAT CONDUCTION, TURBULENCE131


let be the size of an eddy and v the typical rotational velocity
across the eddy; let further  be the energy per unit mass in such
an eddy; then, the energy flow through an eddy of that size is
!
!1
v3
v2


(10.95)

2
v

|{z}
|{z}
typical energy time scale
energy is being fed into the turbulent cascade on the macroscopic
scale L where the typical velocity is u; from there, the energy cascades through the turbulent eddies to progressively smaller scales
until it is finally viscously dissipated on a scale v ; in between,
i.e. on scales satisfying
s < < L ,

(10.96)

the energy flow  must be independent of scale because the energy


cannot be accumulated anywhere; therefore, we conclude from
(10.95) that the typical eddy velocity must change with the eddy
scale as
v 1/3
(10.97)
or
v u

 1/3

;
(10.98)
L
the largest eddies thus carry the highest velocities, but the smallest
have the highest vorticity,

v
u
2 1/3 ;

( L)

(10.99)

to estimate the scale v , we compare the viscous dissipation with


the specific energy flow  ; the viscous heating rate is approximately
hv
2

 v 2

v3

!2/3
4/3 =  2/3 4/3 ;

(10.100)

therefore, hv is negligibly small on large scales, but if the heating


rate becomes of order the energy flow rate,
hv  ,

(10.101)

viscous dissipation sets in; this happens on a length scale v given


by
!3/4

2/3 4/3
 v  v =
(10.102)
 1/3
or, because of


u3
,
L

(10.103)

CHAPTER 10. INSTABILITIES, CONVECTION, HEAT CONDUCTION, TURBULENCE132


the viscous scale can be transformed to
!3/4
 3/4
L1/3
L
=L
v =
= 3/4 ,
u
uL
R

(10.104)

where R is the Reynolds number on the scale L;


finally, we consider the correlation function of the eddy velocity
v , or rather its Fourier transform, the power spectrum; since
v ( )1/3 ,

(10.105)

the correlation function scales as


v ( )2/3 ,
while its Fourier transform Pv will then scale as

2/3
Pv 3 v k3  k1
 2/3 k11/3 ;

(10.106)

(10.107)

the power per logarithmic k interval will thus scale as


k2 Pv  2/3 k5/3 ,
which is the Kolmogorov turbulence spectrum;

(10.108)

Chapter 11
Collision-Less Plasmas
11.1

Basic Concepts

11.1.1

Shielding; the Debye length

we would like to describe plasmas as fluids; this requires that


collisions be random and short-ranged, such that equilibrium can
be locally established sufficiently fast; the fundamental difference
between plasma physics and the hydrodynamics of neutral fluids
is the long-ranged Coulomb interaction between the particles;
a plasma consists of electrons and ions of charge Ze; the existence
of two different types of charge allows shielding on a characteristic length scale; to calculate it, we put into a plasma with charge
density en a point charge q; Poissons equation for the electrostatic potential then reads
~ 2 = 4e(n+ n ) + 4qD (~x)

(11.1)

if the charge is placed at the coordinate origin;


in thermal equilibrium, the charge densities are
!
e
n exp
,
kT

(11.2)

so that we can write


"
!
!#
e
e
n+ n = n exp
exp
,
kT
kT

(11.3)

with the mean particle number density n ;


sufficiently far from the central charge, the argument of the exponential
e
1
(11.4)
kT
133

further reading: Shu, The


Physics of Astrophysics, Vol II:
Gas Dynamics, chapters 28 and
29; Padmanabhan, Theoretical
Astrophysics, Vol. I: Astrophysical Processes, sections 9.1
9.4; Ishimaru, Basic Principles
of Plasma Physics, chapters 1
4

CHAPTER 11. COLLISION-LESS PLASMAS

134

so that we can approximate


2e
n+ n n
kT

!
(11.5)

and write Poissons equation as


2
~ 2 8e n + 4qD (~x) ,

kT

(11.6)

which has the solution


2r
q
,
= exp
r
D

(11.7)

which contains the Debye length defined by


kT
D
4e2 n

!1/2

 T 1/2  n 1/2
= 6.9 cm
;
K
cm3

(11.8)

this is the characteristic length scale for the shielding of a charge


in a plasma;
a plasma can be considered ideal if it contains sufficiently many
particles within the Debye length, since then the interaction energy is small compared to the thermal energy; to see that, we
compare
e2
with kT ;
(11.9)
r
the first expression is the mean potential energy of the interacting
charges, the second is a measure for their thermal energy; r is the
mean separation of the particles, given by
4 3
r n 1
3

3
r
4n

!1/3
;

(11.10)

we thus obtain from the comparison between the two energies


e2
kT

4n
3

!1/3

4ne2 3
=
3kT 4n

!2/3

1 r2
=
1
3 2D

(11.11)

for D  r;

11.1.2

The plasma frequency

the mean thermal velocity of the electrons is given by


hv2 i

kT
,
me

(11.12)

CHAPTER 11. COLLISION-LESS PLASMAS

135

which means that an electron passes the Debye length in the time
r
r
D
kT me
me

;
(11.13)
tD p
=
4ne2 kT
4ne2
hv2 i
this is the time sale on which the thermal motion of the electrons
can compensate charge displacements by shielding;
the time tD can be transformed into a characteristic frequency for
plasma oscillations,
s
 n 1/2
4ne2
1
=
5.6 104 Hz
,
(11.14)
p
tD
me
cm3
which is the frequency with which charge inhomogeneities can
oscillate against each other; with D and p , we now have two
essential parameters for describing plasmas at hand;

11.2

The Dielectric Tensor

11.2.1

Polarisation and dielectric displacement

external electric fields E~ polarise media, i.e. they induce in these


media a charge displacement which can be described by a po~ is defined by
larised charge density pol ; the polarisation P
~ P
~ = pol ;

(11.15)

the Maxwell equation in vacuum,


~ E~ = 4 ,

(11.16)

changes to
~ E~ = 4( + pol ) = 4 4
~ P
~;

(11.17)

~ E~ + 4P,
~ is defined as an auxilthe dielectric displacement, D
iary field which satisfies
~ D
~ = 4 ;

(11.18)

due to charge conservation, the charge density must satisfy


pol
~ ~jpol = 0 ;
+
t
if we substitute (11.15) here, we find

~
~
P

P
~

+ ~jpol = 0 or ~jpol =
;
t
t

(11.19)

(11.20)

CHAPTER 11. COLLISION-LESS PLASMAS

136

the Maxwell equation

~ 4
P
1 E~ 4 ~ ~  1 E~
~
~
j + jpol =
B=
+
+ 4 + ~j (11.21)
c t
c
c t
t
c
can then be written as
~
~ B
~ = 1 D + 4 ~j

c t
c

(11.22)

where ~j is the external (unpolarised) current;

11.2.2

Structure of the dielectric tensor

we now need a relation between the external field E~ and the po~ for sufficiently weak fields, we assume this relation
larisation P;
to be linear and write for the Fourier-transformed quantities
D i (, ~k) =  ij (, ~k) E j (, ~k) ,

(11.23)

defining the dielectric tensor  ij ;


due to the Fourier convolution theorem, the multiplication (11.23)
in Fourier space corresponds to a convolution in real space; this
~ x, t) can also be influenced by fields at earlier times
means that D(~
~ x ~x, t t);
and other places, E(~
since the fields must remain real in real space, the dielectric tensor
must satisfy the symmetry relation
 ij (, ~k) =  ij (, ~k)

(11.24)

in Fourier space;
the principal-axis directions of the tensor  ij can only depend on
the vector ~k, so that we can start from the ansatz
 ij (, ~k)

ij +
= A

ki k j
B
k2

with functions A and B depending on (, ~k); obviously,


!
ki k j j
k = ki
k2

(11.25)

(11.26)

is parallel to ~k, while


ij

!
ki k j j
2 k = ki ki = 0
k

(11.27)

CHAPTER 11. COLLISION-LESS PLASMAS

137

and thus perpendicular to ~k; therefore, we split the tensor  ij into


two parts,
!
ki k j
ki k j
i
i
 j = j 2 t + 2 l ,
(11.28)
k
k
which we call the transversal and the longitudinal components of
 ij , with A = t and B = l t according to the ansatz (11.25);
of course, t and l are generally functions of and k which also
need to satisfy the symmetry condition (11.24),
l,t (, k) = l,t (, k) ;

(11.29)

we have neglected in this derivation of the form of the dielectric


tensor  ij that preferred macroscopic directions may exist in the
plasma, e.g. due to magnetic fields ordered on large scales; if they
exist, they must also be built into the dielectric tensor;
in the limit of long waves, |~k| 0, the dielectric tensor tends to
 ij ij t ,

(11.30)

and thus turns into the normal scalar dielectricity;

11.3

Dispersion Relations

11.3.1

General form of the dispersion relations

the dielectric tensor determines which kinds of wave can propagate through the plasma; we shall now derive the dispersion relations between the wave vectors ~k and the frequencies of such
possible waves;
if the fields are decomposed into Fourier modes exp[i(~k~xt)],
Maxwells equations in the plasma read
~
~
~k E~ = B , ~k B
~ = D ,
c
c

~k D
~ = 0 , ~k B
~=0,

(11.31)

if we neglect external charge densities and currents;


combining the curl of the first equation (11.31) with the second
yields
2
~k (~k E)
~ = ~k B
~ = D
~ ,
(11.32)
2
c
c

CHAPTER 11. COLLISION-LESS PLASMAS

138

and if we expand the double vector product, we can write the


displacement vector as
2 ~

~ ;
D = k2 E~ ~k(~k E)
2
c

(11.33)

we now use the dielectric tensor to write D i =  ij E j and find,


written in components
2 i j
 E = k j k j E i ki k j E j ,
c2 j
or
ij

!
ki k j
2 i j
2 2 2  j E = 0 ;
k
ck

(11.34)

(11.35)

the expression in brackets can be represented by a matrix; equa


tion (11.35) has non-trivial solutions E~ , 0 if and only if the
determinant of that matrix vanishes,
!
ki k j
2 i
i
(11.36)
det j 2 2 2  j = 0
k
ck

11.3.2

Transversal and longitudinal waves

this condition defines a relation between ~k and which must be


satisfied for waves allowed to propagate through the plasma; we
now insert  ij from (11.28) and find
(
"
!
#)
ki k j
ki k j
ki k j
2
i
i
det j 2 2 2 j 2 t + 2 l = 0 , (11.37)
k
ck
k
k
or
"
det

ij

ki k j
2
k

!
#
2
2 ki k j
1 t 2 2 2 4 l = 0 ;
ck
c k

(11.38)

for transversal waves, ki E i = 0, and the corresponding terms disappear from the matrix equation (11.35); what remains is
!
2 i
(11.39)
1 t 2 2 E = 0 ,
ck
which implies the dispersion relation
2 =

c2 k2
;
t

if, conversely, E~ is parallel to ~k, terms proportional to


!
ki k j
i
j 2
k

(11.40)

(11.41)

CHAPTER 11. COLLISION-LESS PLASMAS

139

do not contribute, and the dispersion relation is reduced to


2 ki k j E j
l = 0 ,
c2 k4

(11.42)

which generally demands that l = 0; in order to understand this


condition, we need to determine the form of l ;

11.4

Longitudinal Waves

11.4.1

The longitudinal dielectricity

in order to determine the form of l , we return to the collision-less


Boltzmann equation; we neglect the motion of the ions because
of their lower velocities and write
f = f0 + f ,

(11.43)

~ will change
i.e. we expect that sufficiently weak fields E~ and B
the phase-space distribution function only little away from a homogeneous and isotropic distribution function f0 ; to first approximation, Boltzmanns equation then reads
!
~
f
v
~ f e E~ + B
~ f0 = 0 ;
(11.44)
+ ~v
t
c
~p
for an isotropic distribution f0 , we must further have
f0
k ~v
~p

(11.45)

because no other direction is defined, thus


~ f0 = 0 ,
(~v B)
~p

(11.46)

and Boltzmanns equation in linear approximation shrinks to


f
~ f = eE~ f0 ;
+ ~v
t
~p

(11.47)

decomposing E~ and the perturbation f into plane waves


exp[i(~k ~x t)], (11.47) yields
f0
i f + i~v ~k f = eE~
,
~p

(11.48)

which can be solved for f ,


f =

eE~

f0
;
i(~k ~v ) ~p

(11.49)

CHAPTER 11. COLLISION-LESS PLASMAS

140

charge and current densities are exclusively caused by f because


f0 is a homogeneous, isotropic and stationary equilibrium distribution; therefore,
Z
Z
3
~
= e d p f , j = e d3 p~v f ;
(11.50)
these quantities are then also proportional to the phase factor
exp[i(~k~xt)], and the polarisation equations (11.15) and (11.20)
can be written
~ = , iP
~ = ~j ;
i~k P
(11.51)
with f from (11.49), we find
Z
1

f0
2
~
~
ik P = e
d3 p E~
;
~p i(~k ~v )

(11.52)

this integral has a pole at = ~k ~v and is therefore ill-defined; we


remedy that by writing
1
i(~k ~v )

1
i(~k ~v i)

(11.53)

which has no poles any more on the real axis, and later sending
to zero;
~ = D
~ E~ = (l 1)E~ and
for longitudinal waves, we have 4P

E~ = E~k/k; inserting this into (11.52) allows us to write the longitudinal part l of the dielectricity as
Z
1
4e2
f0
l = 1 2
d3 p ~k
;
(11.54)
k
~p (~k ~v i)
if we now place the coordinate system such that ~k points into the
positive ~x direction, the integral can be split up; we then have
~k f0 = k d f0
dp x
~p

(11.55)

and
f0
1
d3 p~k
~p i(~k ~v i)
Z
1
d f(p x )
= k
dp x
,
dp x i(kv x i)
Z

(11.56)

with the definition


Z
f(p x )

dpy dpz f0 (~p) ;

(11.57)

CHAPTER 11. COLLISION-LESS PLASMAS

11.4.2

141

Landau Damping

obviously, the longitudinal dielectricity l has a real and an imaginary part; the latter implies dissipation of electrical energy, as we
shall shortly see; to begin with, the dissipation follows from

E~ 2 ~ ~ E~ E~ ~ ~
+ E j =
Q =
+EP
t 8
4
 E~ D
~
E~  ~

~
,
=
E + 4P =
4
4

(11.58)

~ are to be taken as
where the missing hats indicate that E~ and D
functions of ~x and t here;
we consider the contribution of individual Fourier modes (, ~k) to
the dissipation Q, i.e. we set
~
E~ = E~ ei(k~xt) ,

~ =D
~ ei(~k~xt) ,
D

(11.59)

and thus Q will be the dissipation per Fourier mode (, ~k);


in order to arrive at a real quantity, we need to replace E~ by
(E~ + E~ )/2 to obtain the real part of the E~ field which is defined
~
complex here, and the same for D;
~ = l E,
~ we can write
since, for longitudinal waves, D


~ = i l E~ l E~ ,
D
2

(11.60)

where the minus sign on the second term comes from the change
in sign in the phase factor exp[i(~k ~x t)] due to the complex
conjugation of the E~ field; inserting this into Q from (11.58) gives
Q=


i  ~ ~   ~
E + E l E l E~ ;
16

(11.61)

averaging this over time eliminates the products E~ E~ and E~ E~


because they vary with the phase factor exp(2it), while the
mixed terms become independent of time; thus,

 ~ 2
i  ~ ~
i

Q=
l E E l E~ E~ =
l l |E|
; (11.62)
16
16
the remaining expression in brackets is
l l = 2i=l ,

(11.63)

~ 2 ;
=l |E|
8

(11.64)

and so we find
Q=

CHAPTER 11. COLLISION-LESS PLASMAS

142

the imaginary part of l follows from (11.54),


~k ~v + i

=
,
~k ~v i
(~k ~v )2 + 2 (~k ~v )2 + 2
(11.65)
and in the limit 0 this turns into a Dirac delta function,
=

==

(~k ~v )2 + 2

D (~k ~v ) ;

therefore, the imaginary part of l is


Z
d f
42 e2
dp x k
D (kv x )
=l = 2
k
dp x

42 e2 m d f
=

k2
dp x p =m/k

(11.66)

(11.67)

the damping then turns into



2
2 me d f
~
Q = |E|
;

k2 dp x px =m/k

(11.68)

this is Landau damping, which is caused by the fact that electrons


which are slightly faster than the wave are slowed down, electrons which are slightly slower than the wave are accelerated, and
since the velocity distribution is typically monotonically decreasing, more electrons need to be accelerated than decelerated, and
thus the wave loses energy;

11.5

Waves in a Thermal Plasma

11.5.1

Longitudinal and transversal dielectricities

in a thermal plasma, f0 is the Maxwell distribution


!
n
p2
dp x dpy dpz ,
f0 =
exp
(2mkT )3/2
2mkT
therefore the integrated distribution f is
!
2
p2x /(2mkT )
p
n

(2mkT
)
n

e
x
f(p x ) =
exp
=
,
(2mkT )3/2
2mkT
(2mkT )1/2

(11.69)

(11.70)

and its derivative is


n p x epx /(2mkT )
d f
=
;
dp x
2(mkT )3/2
2

(11.71)

CHAPTER 11. COLLISION-LESS PLASMAS

143

with these expressions, the longitudinal dielectricity can be transformed to read


Z
2
epx /(2mkT )
4e2
n p x
dp x
l = 1 +
; (11.72)
k

2(mkT )3/2 kpmx i


introducing the Debye length D (11.8) and using v x = p x /m, we
find
Z
2
1
mv x
emvx /(2kT )
l = 1 + 2
dv x
,
(11.73)
kD
2mkT kv x i
and with the mean thermal electron velocity
r
kT
vx
,
and z
ve =
m
ve

(11.74)

we get the expression


Z
2
1
zez
1
l = 1+
dz
1+ 2 2 W(x) , (11.75)
k D
2(kD )2 z x iy
where the abbreviations
x

kve

and y

kve

(11.76)

were used and the function


1
W(x)
2

z ez /2
dz
zx

(11.77)

was defined for x C;


this function W(x) is well-defined and analytic for x in the upper
half of the complex plane, i.e. for =x > 0, and it can analytically
be continued into the lower half-plane by integrating over a closed
(rectifiable and positively oriented Jordan) curve C in C whose
interiour encloses x;
we thus choose the contour such that it runs along the real axis
from R to R, possibly including x in a small extension if =x 0,
and closing along a half-circle in the upper complex plane from
R to R; because of the steep exponential drop of the integrand,
the integral along that half-circle will not contribute in the limit
R , and the integral along the closed contour C will equal
the integral along the real axis,
I
Z
2
2
z ez /2
z ez /2
dz

dz
,
(11.78)
zx
zx
C

with =z = 0, as required for evaluating (11.75);

CHAPTER 11. COLLISION-LESS PLASMAS

144

beginning with =x > 0, we can integrate over z along the real


axis; for doing so, we first substitute the identity
Z
1
=i
dt exp [i(z x)t]
(11.79)
zx
0
into (11.77), which yields
W(x) =
=

Z Z
i
2
dt
dz zez /2+i(zx)t

2 Z0
Z

i
2
ixt
dt e
dz zez /2+izt ; (11.80)

2 0

carrying out the z integration first, and splitting the remaining exponential exp(ixt) into trigonometric functions, yields
Z
2
W(x) =
dt t (cos xt i sin xt) et /2
0
r
Z x
x2 /2
y2 /2
x2 /2
xe
; (11.81)
= 1 xe
dy e
+i
2
0
series expansions of W(x) are useful for practical calculations; for
small x, |x| < 1,
r
r
x4
x2 /2

x4
2
2
W(x) 1 x + + i
xe
1 x + +i
x (11.82)
3
2
3
2
while, for large x,
3
1
W(x) 2 4 + i
x
x

x2 /2
xe
;
2

(11.83)

substituting this back into (11.75) and expanding the abbreviations (11.76), we find the longitudinal dielectricity
"
1
(kve )2 3(kve )4
l = 1
+
(kD )2 2
4
#

2 /(2k2 v2e )
i
e
2kve
!
 2
3(kve )2
p
1
1+

2p 2 /(2k2 v2 )
e ;
e
+ i
(11.84)
3
2(kve )
in the limiting case  kve ; we have inserted the plasma frequency p = ve /D in the first term here;

CHAPTER 11. COLLISION-LESS PLASMAS

145

in the opposite limiting case  kve ,


"
!#

2
1
1
l = 1 +
+i
1 2 2
(kD )2
(kve )2
2k ve
2kve

!2
!2


1
;
+i
(11.85)
1+
kve
kve
2kve
a similar calculation leads to the transversal dielectricity
"
!
#
2p

t = 1 + 2 W
1 ;
(11.86)

kve
if ions need to be taken into account in addition to the electrons,
the dielectricities are summed according to
X
 1 =
(i 1) ;
(11.87)
i=species

11.5.2

Dispersion Measure and Damping

the dispersion relation for transversal waves was given by (11.40);


in the high-frequency limit,
t 1

2p
2

(11.88)

because then x  1 and F(x) 1, and thus the dispersion


relation becomes
2 2p = k2 c2

2 = 2p + k2 c2 ;

the group velocity of such transversal waves is


q
s
2 2 c

2
p
2p
kc
cg =
=
=
=c 1 2 ;

(11.89)

(11.90)

the propagation time of such waves is this

Z
Z
Z
2p L 2e2

dl

t =
dl n , (11.91)

dlc 1 +
= +
cg
22
c mc2
where the quantity
Z
dl n DM

(11.92)

is called the dispersion measure (which is the column density of


the free electrons); here, we have approximated  p , which
should not be confused with the assumption  kve !

CHAPTER 11. COLLISION-LESS PLASMAS


for < p , the dispersion relation implies
q
i 2p 2
,
k=
c

146

(11.93)

i.e. the transversal waves are damped;  kve and < p =


ve /D are possible if both ve and D are small, which is the case
in sufficiently cold plasmas; in the ionosphere of the Earth, n
106 cm3 , thus
D 0.11 cm
(11.94)
and
ve 6.4 106 cm s1 ,

(11.95)

p 60 MHz ,

(11.96)

thus
which corresponds to a wavelength of

c
5m ,
p

(11.97)

where all numbers are given assuming T = 273 K;


the dispersion relation for longitudinal waves requires l = 0, as
was shown above [see (11.42)]; to lowest order in kve /,
l 1

2p
2

= p ;

(11.98)

to next higher order, setting the real part of l to zero,


2p

"
#
3(kve )2 !
<l 1 2 1 +
=0,

2
implies

2p

34p k2 2D

+
1=0;
2
4
this quadradatic equation in 2 has the solutions
2 =

2p 
2

q

1 + 12k2 2D

(11.99)

(11.100)

(11.101)

or


p 1 + 3k2 2D ;

(11.102)

Chapter 12
Magneto-Hydrodynamics
12.1

The Magneto-Hydrodynamic Equations

12.1.1

Assumptions

magneto-hydrodynamics is built upon several assumptions which


go significantly beyond hydrodynamics; this begins the fact that
plasmas consist of ions and electrons which should be described
as two fluids which are coupled to each other; this is simplified
by the assumption that ions and electrons may be coupled to each
other so tightly by the electrostatic attraction that they can be
treated as a single fluid;
secondly, the usual assumption of hydrodynamics plays a critical
role that the mean free path of electrons and ions is very small
compared to any other relevant scales;
thirdly, there is usually a small drift velocity ~vdrift between the
electrons and ions,
vdrift = ~ve ~vi ,
(12.1)
which causes an electric current, which in turn induces a magnetic
field;
finally, it is assumed that the plasma flows non-relativistically,
such that terms of higher than linear order in v/c can be neglected;
we thus get to the following mathematical assumptions: in the rest
frame of the plasma, the charge density 0 be negligible compared
to the current density ~j0 ,
0 

147

|~j0 |
;
c

(12.2)

further reading: Shu, The


Physics of Astrophysics, Vol II:
Gas Dynamics, chapter 21 and
27; Padmanabhan, Theoretical
Astrophysics, Vol. I: Astrophysical Processes, section 9.6

CHAPTER 12. MAGNETO-HYDRODYNAMICS

148

the current density be related to the electric field E~ 0 through


Ohms law,
~j0 = E~ 0 ,
(12.3)
with the conductivity , which is assumed to be very high
(ideal), such that weak fields can be responsible for significant
currents;
because of the non-relativistic velocities, we neglect the displacement current,
1 E~ 0
4~j0

;
(12.4)
c t
c
the equations of the magnetic field thus read
~ B
~0 = 0 ,

~ B
~ 0 = 4 ~j0 ;

(12.5)

the rest frame of the plasma and the observers laboratory frame
are related by the Lorentz transformation, which can be written,
to lowest order in v/c, as
0 =

~j ~v
,
c2

~j0 = ~j ~v ;

(12.6)

since we have assimed 0  ~j0 /c, we also have  ~j/c, and


because
~ E~ 0 = 40 ,
~ B
~ 0 = 4 ~j0 ,

(12.7)
c
~  | B|;
~
this also implies |E|
thus, the assumptions of magneto-hydrodynamics lead to the conditions

~
~
|~j|
~  | B|
~ , E  j , |~v|  1 ;
, |E|
(12.8)

t c
c
c
from here, we can now derive the equations of magnetohydrodynamics;

12.1.2

The induction equation

from Maxwells equation


~
1 B
~ E~
=
c t

(12.9)

~
v

~j ~j0 = E~ 0 = E~ +
,
c

(12.10)

and Ohms law

CHAPTER 12. MAGNETO-HYDRODYNAMICS


we conclude

and

149

~j ~v B
~
E~ =

(12.11)

~
~
~j ~v B
B
,
~
= c
t

(12.12)

but we also need to satisfy


~ B
~j = c
~,
4

(12.13)

and thus
!
~
B
c2 ~ ~
~
~ (~v B)
~
=
B +
t
4
2 h
i
~
~ B)
~ 2B
~ (~v B)
~
~
~ c (
=
4
2
~ (~v B)
~ 2B
~,
~ + c
=
4

(12.14)

~ B
~ = 0; this induction equation determines the evobecause of
lution of the magnetic field embedded into a plasma flow with the
velocity ~v; moreover, we have assumed that is spatially con~ = 0;
stant,

12.1.3

Eulers equation

we now need equations for the back reaction of the magnetic field
on the flow of the plasma; first, there is the continuity equation
for the mass density , which is of course unalteredly valid,
~
+ (~v) = 0 ;
t

(12.15)

Eulers equation which describes the conservation of momentum or, more precisely, the transport of the specific momentum
density, must be modified because of the Lorentz equation; the
Lorentz force on a charge e is
e
~ ;
(~v B)
c

(12.16)

multiplying that with the number density of charges, the product


e~v turns into the current density ~j, which is
~ B
~j = c
~,
4

(12.17)

and thus Eulers equation is modified to read

d~v
~ + 1 (
~ B)
~ B
~;
= P
dt
4

(12.18)

CHAPTER 12. MAGNETO-HYDRODYNAMICS

150

the Lorentz force density on the right-hand side can also be rewritten,
!
1 ~ ~
1 i jl Bm k
~
B
( B) B =
 
4
4 jk m xl
 Bm
1  il
km im lk Bk l
=
4
! x
i
1
B
B
k
=
Bk k Bk
4
x
xi
1 ~ ~2
1 ~ ~ ~
=
( B ) B
B ;
(12.19)
4
8
~ changes along B,
~ i.e. it quantifies
the first term specifies how B
the tension of the magnetic field lines, which obviously tend to
be as straight as possible; the second term is the gradient of the
magnetic energy density and augments the pressure gradient in
Eulers equation;
we had seen in normal, viscous hydrodynamics that Eulers equation can be written in the form
(vi ) T i j
+
=0,
t
x j

(12.20)

with the stress-energy tensor


T i j = vi v j + Pi j i j

(12.21)

and the shear tensor


!
vi v j 2 i j ~
~ ~v ;
+
~v i j
=
x j xi 3
ij

in presence of the magnetic field, this changes to

~ 2 i j
1 i j B
ij
ij
B B ,
T T
4
2

(12.22)

(12.23)

i.e. the stress-energy tensor of the viscous fluid is augmented by


the stress-energy tensor of the magnetic field;
together with an equation of state P = P(), the induction equation, the continuity equation and Eulers equation determine the
flow; these are eight equations for the eight unknowns , P, ~v
~ if the magnetic field is known, the current follows from
and B;
equation (12.17), and the electric field from (12.11);

CHAPTER 12. MAGNETO-HYDRODYNAMICS

12.1.4

151

Energy and entropy

the entropy equation, which read


T

ds
vi ~
~ ),
(T
= i j j +
dt
x

(12.24)

needs to be augmented by the production of Ohmic heat; per unit


time, the induction current ~j0 in the rest frame of the fluid dissipates the energy
2

2
~02 ~2
~ B
~ ,
~j0 E~ 0 = j j = c

162

(12.25)

which must be added to the right-hand side of the entropy equation,


T

2

2
ds
vi ~
~ )+ c
~ B
~ ;
(T

= i j j +
dt
x
162

(12.26)

if we need to express this by the specific energy density instead


of the specific entropy density s, we start from the energy conservation equation of viscous hydrodynamics and augments that
analogously; this yields

~ 2
B
~v2
~ ~q = 0
+ + +
(12.27)

t 2
8
with the extended energy current density vector
!
~v2
T
c
i
+ w vi
ij v j +  ijk E j Bk ,
q =
2
xi
4
which now contains the Poynting vector

~
c ~ ~
c ~j ~v B
B
~
~

S =
EB=
4
4
c
c2 ~ ~
~ 1 (~v B)
~ B
~,
( B) B
=
2
16
4

(12.28)

(12.29)

and this yields the energy current density


!
~v2
T
i
q =
+ w vi
ij v j
(12.30)
2
xi
ii
h
ii
c2 h ~
~ B)
~ + 1 B
~ (~v B)
~ ;

162
4
~ ~v = 0, these equations simplify
for incompressible flows with
somewhat; then, we first have to satisfy
~ B
~=0,

~ ~v = 0 ,

(12.31)

CHAPTER 12. MAGNETO-HYDRODYNAMICS

152

next, the induction equation,


2
~
B
~ (~v B)
~ 2B
~ + c
~
=
t
4

c2 ~ 2 ~
~
~
~
~
~
~
~
~
= ~v( B) B( ~v) + ( B )~v (~v ) B +
B
4
2
~ v (~v )
~ B
~ 2B
~ )~
~+ c
~,
(12.32)
= (B
4
thus

2
~
B
~ B
~ v+ c
~ 2B
~ = (B
~ )~
~,
+ (~v )
t
4
~ ~v = 0
and Eulers equation reads with

(12.33)

~ 2

1
B
~v
~ v=
~ P + + 1 ( B
~ B
~ 2~v , (12.34)
~ )
~ +
+ (~v )~
t

8
4
where = / is the specific viscosity per unit mass; likewise, the
viscosity tensor ij in the energy conservation equation simplifies;

12.1.5

Magnetic advection and diffusion

two terms determine the temporal change of the magnetic field in


the induction equation,
~ (~v B)
~

and

c2 ~ 2 ~
B;
4

(12.35)

~ (~v B),
~ determines the transport of the magnetic
the first term,
field with the fluid flow; it is called advection term; its order-ofmagnitude is
vB
,
(12.36)
L
if L is a typical length scale of the flow;
~ 2 B/(4),
~
the second term, c2
determines the diffusion of the
magnetic field due to the finite conductivity of the plasma; its
magnitude is of order
c2 B
,
(12.37)
4 L2
and it vanishes if the conductivity is ideally (infinitely) large;
the ratio of the two orders-of-magnitude,
advection 4 L2 vB 4vL
= 2
=
RM ,
diffusion
c B L
c2

(12.38)

CHAPTER 12. MAGNETO-HYDRODYNAMICS

153

is called the magnetic Reynolds number; obviously, the diffusion


can be neglected if RM  1, and the induction equation simplifies
to
~
B
~ (~v B)
~ =0

(12.39)
t
in this case, there is no diffusion, and the magnetic field is
frozen into the plasma; the physical reason for that is that, if
the conductivity is ideally high, , each motion of the magnetic field with respect to the plasma immediately induces strong
currents which counter-act their origin, i.e. the motion of the field;
this is the typical case in astrophysical plasmas;
in the opposite limit, RM  1, which occurs of the conductivity
is small, the induction equation reads
~
B
c2 ~ 2 ~
=
B;
t
4

(12.40)

the magnetic diffusion time scale therefore is


diff 4

L2
;
c2

(12.41)

plasmas in the laboratory are typically characterised by RM  1,


while astrophysical plasmas typically have RM  1;

12.2

Generation of Magnetic Fields

the induction equation constains no source term, i.e. it only de~ = 0 initially,
scribes how existing magnetic fields change, but if B
this remains so; this is a consequence of the assumption that ions
and electrons are ideally (tightly) coupled to each other;
if that is not the case, the motions of the electrons and the ions
need to be considered separately, in particular with different velocities ~ve and ~vi ; then, the two separate Euler equations for the
electrons and the ions are
!
~
d~ve
v
e
~ e ne e E~ + B
~ ,
~ ne me
ne me
= P
dt
c
!
~
d~vi
v
i
~ i ni e E~ + B
~ ; (12.42)
~ ni mi
ni mi
= P
dt
c
we divide these equations by ne me and ni mi and subtract the sec-

CHAPTER 12. MAGNETO-HYDRODYNAMICS

154

ond from the first; this yields


~ e
~ i
d(~ve ~vi )
P
P
=
+
dt
ne me ni mi
!
e ~ ~ve ~

E+ B
me
c
!
e ~ ~vi ~

E+ B ;
mi
c

(12.43)

since the ion mass mi is much larger than the electron mass me ,
but ne = ni n, equation (12.43) can be approximated by
!
~ e
d(~ve ~vi )
P
e ~ ~ve ~
=

E+ B ;
(12.44)
dt
nme me
c
this equation must be accomplished by a phenomenological collision term through which different electron and ion velocities can
be justified or produced in the first place; introducing a collision
time , we write
!
~ e
~ve ~vi
d(~ve ~vi )
P
e ~ ~ve ~
=

E+ B
;
(12.45)
dt
nme me
c
dt
the current is
~j = eni~vi ene~ve = en(~vi ~ve ) ,

(12.46)

where we emphasise the implicit assumption of singly-charged


ions;
we now assume that the relative drift velocity is constant,
d(~vi ~ve )
=0,
dt

(12.47)

which also means that the current density is constant; in such a


situation, the electric field following from (12.44) is
~ e ~ve
P
~
E~ =
B
ene
c
~ e ~ve
P
~+
=
B
ene
c

me
(~vi ~ve )
e
me ~j
;
ne2

(12.48)

this implies

~ e ~ve
~
~j
P
B
m
e
~ E~ = c
~
~

= c
en + c B
t
ne2
e
=

~ e
~
c ~ P
~ (~ve B)
~ j ; (12.49)
~ me c

+
2
e
ne
e
n

CHAPTER 12. MAGNETO-HYDRODYNAMICS


now, we have

~
~
~ Pe = P
~ e n

ne
n2

155

(12.50)

~ P
~ e vanishes identically, and
because
~ ~
~
~
~ j = ~j n + j

n
n2
n
~
c ~
~ B)
~ + ~j n ;
=
(
4n
n2

(12.51)

if the current is flowing along the gradient in the number density


of the electrons, the latter term vanishes identically, and we obtain
the modified induction equation
~
B
me c2 ~ 2 ~
c ~
~
~
~

n)
+

(~
v

B)
+
= 2 (P
B;
e
e
t
en
4ne2

(12.52)

inserting the definition

ne2
(12.53)
me
of the conductivity, this equation is identical to the previous form
of the induction equation, except for the first term,
c ~
~ ,
(Pe n)
(12.54)
en2

whcih now appears as the source of the magnetic field; mechanisms like this for creating magnetic fields are called battery
mechanisms;

12.3

Ambipolar Diffusion

12.3.1

Scattering cross section

in a mixture of neutral particles and plasma, the magnetic field


can be thought of frozen into the plasma; collisions between the
plasma and neutral particles then creates a friction force between
the plasma and the neutral fluid, which causes a diffusion of the
magnetic field; this diffusion process is called ambipolar;
in order to determine it, we first need a cross section for the collisions, or, more conveniently, the velocity-averaged cross section
hvi; we adopt two limiting cases for it;
if v is very large, we may approximate the cross section by its
geometrical value; if ri and rn are the effective radii of the ions
and the neutral particles, respectively, we have
= (ri + rn )2

(12.55)

CHAPTER 12. MAGNETO-HYDRODYNAMICS

156

and thus
hvi = hvi = hvi(ri + rn )2 ;

(12.56)

the relative velocity during the interaction is


v = |~vi ~vn | ;

(12.57)

if v is sufficiently small, the ion can polarise the neutral particle,


by which the cross section is enlarged; the electric field of the ion,
assumed singly charged, is
Ze
E~ i = 2 ~er ,
r

(12.58)

while it should be possible to approximate the electric field of the


polarised neutral particle as a dipole field,
2~p
E~ n = 3 ,
r

(12.59)

where ~p = E~ i is the plarised dipole moment with a parameter ;


the dipole field of the neutral particle exerts the force
2 2

Ze
Z e
F~ = ZeE~ n = 2Ze 5 ~er = 2 5 ~er
r
r

(12.60)

on the ion, which has the potential


U=

Z 2 e2
;
2 r4

(12.61)

the motion of the ion past the neutral particle can be characterised
by the minimum separation r0 , the impact parameter b and the
velocity v at infinity;
angular-momentum conservation requires
v b = r0 v0

(12.62)

mi mn
;
mi + mn

(12.63)

with the reduced mass

in addition, energy conservation demands


2

Z 2 e2
;
v = v20
2
2
2r04

(12.64)

from that, and because of


v0 =

bv
,
r

(12.65)

CHAPTER 12. MAGNETO-HYDRODYNAMICS

157

we obtain a quartic equation for the minimum separation,


r04 = b2 r02
or
2
r0,

b2
=

Z 2 e2
,
v2

b4 Z 2 e2

;
4
v2

(12.66)

(12.67)

2
both roots are mathematically possible, but physically, only r0,+
is
relevant because in the limiting case of vanishing coupling , the
minimum radius must equal the impact parameter, r0 = b, since
then the ion is not scattered at all; r0 will become minimal of the
impact parameter is
!1/4
4Z 2 e2
b0 =
;
(12.68)
v2

since the force between the ion and the neutral particle decreases
very steeply with r, by far the strongest effect occurs for close
encounters; thus, we estimate the cross section as
r
2Ze
2
= b0 =
(12.69)
v

obviously, hv i is independent of the asymptotic velocity v ,


r

hv i = 2Ze
;
(12.70)

12.3.2

Friction force; diffusion coefficient

a single collision transfers the momentum


|~p| = |~vi ~ve | ,

(12.71)

and the scattering rate per volume is ni nn hv i, thus the momentum transfer per time and volume is
f~r = ni nn hv i(~vi ~ve ) ,

(12.72)

which corresponds to a friction force;


with
i n = ni mi nn mn = (mi + mn )ni nn ,

(12.73)

this can be cast into the form


f~r = i n (~vi ~vn ) ,

(12.74)

where is the friction coefficient

hv i
;
mi + mn

(12.75)

CHAPTER 12. MAGNETO-HYDRODYNAMICS

158

in the two limiting cases of very small or very large relative velocity, we find
(ri + rn )2
|~vi ~vn | or
mi + mn
r
2Ze

=
mi + mn

(12.76)

if the Lorentz force and the friction force balance each other, the
relative drift velocity between the ions and the neutral particles is
correspondingly established; the Lorentz force density is
~j B
~
1 ~ ~
~,
f~L =
=
( B) B
c
4

(12.77)

from which follows


i n (~vi ~vn ) =
or
~vd ~vi ~vn =
this is of order
vd

1 ~ ~
~;
( B) B
4

(12.78)

~ B)
~ B
~
(
;
4i n

(12.79)

B2
;
4i n L

(12.80)

the magnetic field, which is thought to be frozen into the flow


of the plasma, satisfies the equation
~
B
~ (B
~ ~vi ) = 0 ;
+
t

(12.81)

in order to calculate the flow of the magnetic field relative to the


neutral particles, we replace ~vi by ~vd , which leads to the equation

~ B)
~
~ B
~

B
~ B
+
(12.82)
~ 4 = 0 ;
t
i n
the second term is of order
B3
,
4i n L2

(12.83)

and corresponds to a diffusion coefficient


D

B2
vd L ;
4i n

(12.84)

Chapter 13
Waves in Magnetised Plasmas
13.1

Waves in magnetised cold plasmas

13.1.1

The dielectric tensor

we study the propagation of electromagnetic waves in a plasma in


which random particle motion is negligible, whose temperature is
thus low, and which can thus be considered as cold;
the equation of motion of an electron in such a plasma with ex~ 0 is then only determined by the Lorentz
ternal magnetic field B
force,
~v ~
d~v
(13.1)
m = eE~ e B
0 ,
dt
c
where E~ is the internal electric field;
we now assume that the spatial change of the electric field is negligible and decompose ~v and E~ into harmonic contributions varying with time as exp(it); then, each monochromatic velocity
mode is determined by
!
e ~ ~v ~
(13.2)
i~v = E + B0 ,
m
c
which implies the matrix equation

e i k j e i
i
i j
 B v = E ;
mc jk 0
m

(13.3)

the term in parentheses on the left-hand side can be represented


by the matrix


i 0 0 0
3 2

0 i 0 3 0
1 M ,
(13.4)

1
0 0 i

0
159

further reading: Shu, The


Physics of Astrophysics, Vol
II: Gas Dynamics, chapter 22;
Padmanabhan, Theoretical Astrophysics, Vol. I: Astrophysical Processes, sections 9.7;
Ishimaru, Basic Principles of
Plasma Physics, chapter 5

CHAPTER 13. WAVES IN MAGNETISED PLASMAS

160

where we have identified the Larmor frequency


eB0
mc

(13.5)

Bi
eB0 Bi0
= B 0 ;
mc B0
B0

(13.6)

B
and used the definition
i

the inverse of the matrix (13.4) is


1
M 1 =
det M

21
1 2 i3 1 3 i2

i
2 22
1 3 i2
1 2
3

1 3 i2 1 3 i2
2 23

(13.7)

det M = i(2 2B ) ,

(13.8)

with
which can be written in components as
 j 1


i
j l
2 j
j
Mk
=

i

;
k
k
kl
(2 2B )

(13.9)

this allows us to write the solution for the velocity as


e
~v = M 1 E~ ,
(13.10)
m
which yields, with the inverse matrix (13.9), the velocity components v j


ie
j k l
2 j
j
k

i
E

, (13.11)
vj =
k
kl
m(2 2B )
or, written in vector notation,
"
#
2B
ie
iB ~
~
~
~v =
E 2 (E ~eB )~eB
E ~eB ,

m(2 2B )

(13.12)

~ 0 /B0 is the unit vector in the direction of the external


where ~eB B
~ 0;
magnetic field B
~ = D,
~ the polarisation is
because of E~ + 4P
~ ~
~ = DE ;
P
4

(13.13)

its temporal change is the current density


~
~ ~
~j = ene~v = P = iP
~ = i D E ,
t
4

(13.14)

again under the assumption of a harmonic dependence on time,


eit ;

CHAPTER 13. WAVES IN MAGNETISED PLASMAS

161

therefore, we find for the dielectric displacement


~ = 4i ~j + E~ = E~
D
(13.15)

"
#
4e2 ne ~ 2B ~
iB ~
E 2 (E ~eB )~eB
+
E ~eB ;

m(2 2B )
if we can write this in the form D j = kj E k , we can read off the
dielectric tensor; we thus obtain
"
#
4e2 ne
4e2 ne 2B j
j
j
k = k 1
+
e eBk
m(2 2B )
m(2 2B ) 2 B
4ie2 ne B j l
+
 e ,
(13.16)
m(2 2B ) kl B
which we abbreviate as
kj  kj + (k  )eBj eBk + igklj elB ;
if we identify the plasma frequency there,
r
4e2 ne
p =
,
m

(13.17)

(13.18)

we obtain the expressions


2p

,
2 2B
2p 2B
k =  + 2
2B 2
!
2p
2p
2B
= 1 2
1

=
1

,
2
2
2B
2p B
;
g =
2 2B

 = 1

13.1.2

(13.19)

Contribution by ions

if ions need to be taken into consideration, these quantities change


according to
 1 ( 1)e + ( 1)i ,
k 1 (k 1)e + (k 1)i ,
g ge + gi ;

(13.20)

there, the plasma and Larmor frequencies of the electrons and the
ions will then have to be distinguished;

CHAPTER 13. WAVES IN MAGNETISED PLASMAS

162

the larmor frequency of the ions is


Bi =

ZeB0 Zme
=
Be  Be ;
mi c
mi

(13.21)

the plasma frequency of the ions is given by


2pi =

4Z 2 e2 ni 4Z 2 e2 ne Zme 2
=
=
,
mi
mi
mi pe

therefore its ratio to the electron plasma frequency is


r
pi
Zme
=
1;
pe
mi

(13.22)

(13.23)

thus, the contribution of the ions to the longitudinal dielectricity


k is negligible, but not necessarily their contribution to the perpendicular dielectricity  ; the ratio
2
pi 2 2 2

2 2Bi pi 2Be
(13.24)
2 = 2 2
1
pe 2Bi
2 pe 2
Be
only if

2
2Be 1
, with
2
f 2 2Be f

Zme
mi

(13.25)

f (2 2Be ) 2 f 2 2Be

(13.26)

holds; this is the case if

is satisfied, or
2

f f2 2
Be f (1 f )2Be ;
1+ f

(13.27)

similarly, the contributions of ions and electrons to g are comparably large;

13.1.3

General dispersion relation

with the expressions (13.17) and (13.19), the general dispersion


relation
!
k j kk
2 j
j
det k 2 2 2 k = 0
(13.28)
k
kc
leads to the equation
kc
A

!4

kc
+B

!2
+C = 0 ,

(13.29)

CHAPTER 13. WAVES IN MAGNETISED PLASMAS

163

with the coefficient


ki k j ekB
eB j k j eBk kk
k j kk
+ igi jk 2 ; (13.30)
A  jk 2 =  + (k  )
k
k2
k
if denotes the angle between the magnetic field and the wave,
~eB ~k = cos , this implies
A =  + (k  ) cos2 =  sin2 + k cos2 ,
because

ki k j ek ~k
i jk 2 B =
k
k

(13.31)

~k

~eB = 0 ;
k

(13.32)

B =  k (1 + cos2 ) (2 g2 ) sin2

(13.33)

similarly, we find

and
C = k (2 g2 ) ,

(13.34)

and, by definition, A equals the longitudinal dielectricity l ;

13.1.4

Wave propagation parallel to the magnetic field

~ the angle = 0 and thus


if ~k k B,
A = k ,

B = 2 k ,

C = k (2 g2 )

(13.35)

the roots of the dispersion relation (13.29) are then


!2

q
kc
1 
2 2
2 2
2
=
2 k 4 k 4k ( g )

2k
2p B
2p

=  g = 1 2
2B 2 2B
2p 
B 
= 1 2
1

2B
= 1

2p
( B )

(13.36)

in order to determine what kind of waves are described by these


relations, we return to the equation (11.35)
!
ki k j
2 i j
i
j 2 2 2 j E = 0 ;
(13.37)
k
kc
for longitudinal waves, E~ k ~k, we have
 ij E j

ki k j Ek j
Ek j kk
=  E + (k  ) 2
+ ig ijk
,
k k
k k
i

(13.38)

CHAPTER 13. WAVES IN MAGNETISED PLASMAS

164

~ k ~k; this can obviously be reduced


since we had also assumed B
to
 ij E j =  E i + (k  )E i = k E i ,
(13.39)
and, in addition, we have
!
ki k j Ek j
ki k j j
i
j 2 E = Ei 2
=0;
k
k k

(13.40)

thus, longitudinal waves do not satisfy the dispersion relation


(13.37), which implies that (13.37) describes transversal waves;
for transversal waves, E j k j = 0, and thus
!
k
2
jk
i
i
=0
E 2 2  E + ig jk E
kc
k
i

(13.41)

or, in vector notation,

~k

1
 E~ + ig E~ = 0 ;
E~
 g
k

(13.42)

if we turn the z axis into the direction of ~k, the components E x


and Ey of E~ read

1 
 E x + igEy = 0 and
 g

1 
Ey
 Ey igE x = 0 ,
 g

Ex

(13.43)

which is solved if
E x = iEy ,

(13.44)

characterising circularly polarised light;

13.1.5

Faraday rotation

the two solutions (13.36) for the dispersion relation thus describe
the propagation of left- and right-circular polarised transversal
waves which obey different dispersion relations; left- and rightcircular polarised light thus propagates differently along the magnetic field; the polarisation direction of linearly polarised light is
thus rotated; this effect is called Faraday rotation;
in the limit of  p , we have

2
2 

2
2

p
p

B

1
k2 = 2 1
1
,
2
2
c
( B )
c

(13.45)

CHAPTER 13. WAVES IN MAGNETISED PLASMAS

165

or
k

2p 
2p
2p B
B 

1
=
c
22

c 2c 22 c
k0 k ;
(13.46)

the first term k0 corresponds to the wave vector in the unmagnetised medium, which the second term k causes a phase shift
between left- and right-circular polarised light, and therefore to a
rotation of linear polarisation by the angle
Z
Z 2
Z
p B
4e2 ne eB dz
=
kdz =
dz
=
22 c
m mc 22 c
Z
2e3
dz ne B ;
=
(13.47)
m2 c2 2
obviously, the Faraday rotation is proportional to 2 or, equivalently, to the squared wave length 2 ; the expression
Z
dz ne B RM
(13.48)
is called the rotation measure;

13.1.6

Wave propagation perpendicular to the magnetic field

~ thus = /2, and the coefficients A, B and C


in this case, ~k B,
turn into
A =  ,

B =  k 2 + g2 ,

C = k (2 g2 ) ;

these imply the solutions for the dispersion relation


!2
!
k
C
1
C
kc
=

+
 k

2 2 k 2
k
2
C
g
= k or
= 
,
 k

where
B =  k

C
k

(13.49)

(13.50)

(13.51)

was used;
the first of these solutions is a dispersion relation which is ob~ perpendicular to the magnetic field,
viously independent of B;
wave propagation is thus possible as in an unmagnetised plasma;
~ the
these waves are transversal and polarised in the direction of B;
other dispersion relation corresponds to waves with longitudinal
and transversal components;

CHAPTER 13. WAVES IN MAGNETISED PLASMAS

13.2

Hydro-Magnetic Waves

13.2.1

Linearised perturbation equations

166

now we consider, in a way very similar to the treatment of sound


waves in a neutral fluid, the propagation of waves in a magnetised plasma; for simplicity, we assume that dissipation and heat
conduction are unimportant, = = = 0, and that the conductivity be infinite, 1 = 0; then, the equations of magnetohydrodynamics read
~
B
~ (~v B)
~ (~v) = 0 ,
~ , +
=
t
t
~
~v
~ v = P + 1 (
~ B)
~ B
~;
(13.52)
+ (~v )~
t

~ B
~=0,

the energy conservation equation is not relevant for the following


considerations;
we assume that an equilibrium solution exists,
~0 ,
B

0 ,

P0 ,

~v0 = 0 ,

(13.53)

~v ;

(13.54)

which is perturbed by small quantities


~,
B

P ,

in absence of dissipation, entropy is conserved along flow lines;


we further assume isentropic flow, thus s = const. everywhere in
the flow;
we now linearise the MHD equations and obtain, to first order in
the perturbations
~ B
~=0,

~
B
~ (~v B)
~ ,
=
t

~
+ (~v) = 0 ;
(13.55)
t
with = const. for the homogeneous equilibrium solution, the
last equation implies

~ ~v = 0 ;
+
(13.56)
t
finally, Eulers equation reads to first order in the perturbations
~ + P)
~v
(P
=
t
+
nh
i 
o
1
~ (B
~ + B)
~ B
~ + B
~
+

4( + )
~
~ B)
~ B
~
P
(

+
,
(13.57)

CHAPTER 13. WAVES IN MAGNETISED PLASMAS

167

again under the assumption that the equilbrium solution is homo~ =0=
~ B;
~
geneous, thus P
again, we decompose the perturbations Q into plane waves,
h
i
Q exp i(~k ~x t) ,
(13.58)
and find
~=0,
i~k B

~ = i~k (~v B)
~
i B

(13.59)

for the magnetic field and


i + i~k~v = 0 ,

i~v =

~ B
~
i~kP (i~k B)
+
(13.60)

for the velocity;


the pressure perturbation can be related to the density perturbation
as usual,
P = c2s ,
(13.61)
where cs is the sound speed in the neutral gas, and this allows us
to write Eulers equation as
i~v =

~ B
~
ic2s ~
(i~k B)
k +
;

(13.62)

without loss of generality, we can now rotate the coordinate frame


~ falls into the
such that ~k points along the positive x axis and that B
~
~ with ;
x-y plane; further, we denote the angle between k and B
first, the continuity equation yields
=

~k ~v
kv x
=
;

(13.63)

the phase velocity of the wave is


ck
and thus
=

,
k

(13.64)

v x
;
ck

(13.65)

the induction equation turns into


Bx = 0 ,

By =

v x By vy Bx
,
ck

Bz =

vz Bx
ck

(13.66)

CHAPTER 13. WAVES IN MAGNETISED PLASMAS

168

(the latter because of Bx = 0); and Eulers equation finally reads,


also in components
v x +

kBy By
c2s k
=
,

4
kBx By
vy =
,
4
kBx Bz
vz =
;
4

(13.67)

with (13.64) and (13.65), we obtain from here the two sets of
equations
ck Bz = vz Bx ,

ck vz =

Bx Bz
4

(13.68)

and
Bx By
ck vy =
,
4
!
By By
c2s
c2s
=
ck v x v x = v x ck
;
(13.69)
ck
ck
4
ck By = v x By vy Bx ,

the first set (13.68) of equations contains only Bz and vz ; the


second set (13.69) couples v x and Bx to vy and By ; we can
thus distinguish waves which are purely transversal and are described by (13.68), and other waves, which have longitudinal
and transversal components; since the density perturbations are
caused by velocity perturbations v x , only longitudinal waves are
responsible for them;

13.2.2

Alfven waves

the first set (13.68) of equations, describing transversal waves,


yields
ck vz =

vz B2x
4ck

ck =

Bx
= p
k
4

(13.70)

if we eliminate Bz between the first and the second equation;


~ ~k, this can also be written as
since Bx = B
~ ~k
B
= p
cA k cos ,
4

(13.71)

where the Alfven velocity


B2
cA =
4
was identified;

!1/2
(13.72)

CHAPTER 13. WAVES IN MAGNETISED PLASMAS

169

the phase velocity of Alfven waves is


cA cos ,

(13.73)

~
B

;
= p
4
~k

(13.74)

their group velocity is

accordingly, Alfven waves travel along ~k with the phase velocity


~ furthercA cos , which depends on the angle between ~k and B;
more, they are transversal and transport physical quantities (en~ independent of ~k!
ergy and such) along the magnetic field B,

13.2.3

Slow and fast hydro-magnetic waves

we now set vz = 0 and consider waves which obey the second


set (13.69) of equations; if we eliminate
By =

v x By vy Bx
ck

(13.75)

from the second and third equations, we find



Bx 
v x By vy Bx ,
4

By 
=
v x By vy Bx ;
4

c2k vy =


c2k c2s v x

(13.76)

further eliminating
Bx By v x 2
B2
vy =
ck x
4
4

!1
=

Bx By v x
4c2k B2x

(13.77)

from the second equation yields




B2x B2y v x
B2y v x
+
;
c2k c2s v x =
4
4(4c2k B2x )

(13.78)

v x cancels on both sides, and we find


B2y

!
B2x
1+
4
4c2k B2x
B2y
c2k
=

4 c2k B2x /(4)


!


c2k B2y
B2x
2
2
2
ck cs ck
=
(13.79)
;
4
4


c2k c2s =

CHAPTER 13. WAVES IN MAGNETISED PLASMAS

170

this quadratic equation in c2k has the solutions


c2k,

s
!2
c2s
c2s
B2 c2
B2
B2
(13.80)
=
+

+
x s
2 8
2 8
4
s
!2
2
2
c2s + c2A
cs + cA
=
c2s c2A cos2 ;

2
2

thus, a fast and a slow wave mode are possible;


~ is (almost) parallel to ~k,
we first consider the case in which B
hence  1 and cos2 1 2 /2; then,
r

2
2


c2s + c2A
c2k =
c2s c2A
2
k
2
 ( c2 or
1 2
2
2
2
s
(13.81)
cs + cA |cs cA | =
=
2
c
2
A
to first order in ; accordingly, the fast wave propagates with the
faster of the sound and the Alfven speeds, the slow wave with the
slower;
in this approximation, By B, and the Alfven mode has
v x =

c3A
c2A

c2s

B
,
B

vy =

Bx By
,
4cA

(13.82)

which can be approximated as


v x 0 ,

vy cA

B
;
B

(13.83)

the acoustic mode has


By By = 0 ,

vy =

c2 B
Bx By
A
;
4cs
cs B

(13.84)

~ ~k, we find
if, finally, cos2 = 0 or B
c2k, =


1 2
cs + c2A c2s + c2A ,
2

thus
ck =

c2s + c2A

for the fast and ck = 0 for the slow MHD wave;

(13.85)

(13.86)

Chapter 14
Jeans Equations and Jeans
Theorem
14.1

Collision-less motion in a gravitational


field

14.1.1

Motion in a gravitational field

particles in a gas or a fluid move almost unaccelerated until they


meet another particle, which forces them to change their state
of motion abruptly; we had based our treatment of hydrodynamics on the assumption that the collisions occur on much smaller
length scales than those that characterise the extent of the entire system, L; in plasma physics, we had seen that the shielding
of charges on length scales D allows a hydrodynamical treatment despite the inifite range of electrostatic interactions, provided there are sufficiently many particles in a volume 3D ;
in these cases, the interactions are effectively extremely shortranged; likewise, we had assumed in our treatment of local thermodynamical equilibrium that the mean free path of the photons
is much smaller than the characteristic dimensions of the system
under consideration;
studying the motion of many point masses such as stars in a gravitational field, we encounter a fundamental change: the forces
between the particles are long-ranged and cannot be shielded; a
single star in a galaxy, for instance, thus experiences not only the
attraction of its nearest neighbours, but of a large fraction of all
stars in the entire galaxy;
let us consider now a two-dimensional system, such as a galactic
disk, which we shall assume to be infinitely extended for now and

171

further reading:
Binney,
Tremaine, Galactic Dynamics,
sections 4.14.4

CHAPTER 14. JEANS EQUATIONS AND JEANS THEOREM 172


in whose centre we assume a star; the disk be randomly covered
by stars such that their mean number density is spatially constant;
in a circular ring of radius r and width dr, we find
dN = 2rdr n

(14.1)

stars, whose gravitational force on the star in the centre is


Gm2
(14.2)
r2
of the mass of all stars is assumed to be the same, for simplicity;
dF = 2rdr n

of course, the directions of all forces cancel in the mean, but the
contribution of arbitrarily distant rings diverges logarithmically,
Z
Z
dr
2
dF = 2Gnm
= 2Gnm2 ln r ;
(14.3)
r
thus, the structure of the entire stellar system is important for the
dynamics of the stars in the gravitational field;
in the spirit of our distinction of microscopic and macroscopic
forces, which we had made when introducing hydrodynamics, the
forces in a system which is dominated by self-gravity are also
macroscopic; therefore, the collision terms, which describe the
interaction on a microscopic scale, can be neglected here at least
to first order of approximation;
thus, we begin our treatment of self-gravitating systems with the
collision-less Boltzmann equation,
d f (~x,~v, t) f ~
f
=0;
=
+ ~x f + ~v
dt
t
~v

14.1.2

(14.4)

The relaxation time scale

before we turn to a detailed study of Eq. (14.4) in a gravitational


field, we investigate approximately how the trajectory of a star
through a galaxy which is composed of individual stars deviates
from the trajectory through a hypothetical, smooth galaxy;
we consider the passage of a star by another star employing
Borns approximation, i.e. we integrate the deflection along a
straight trajectory passing the deflecting star at an impact parameter b; the perpendicular force at the location x along the hypothetical, straight trajectory is


2

Gm
Gm2 b
~ | =
F = |
= 2
,
(14.5)

b b2 + x2 (b + x2 )3/2
where is the gravitational potential;

CHAPTER 14. JEANS EQUATIONS AND JEANS THEOREM 173


with x vt, we have
"
 vt 2 #3/2
Gm2
F 2 1 +
,
b
b

(14.6)

and Newtons second law mv = F thus implies


Z "
 vt 2 #3/2
Gm
1+
dt
v
b2
b
Z
2Gm
2Gm
=
(1 + 2 )3/2 d =
;
bv 0
bv

(14.7)

let N be the number of stars in the galaxy and R be its radius, then
the fiducial test star experiences
N = 2bb n = 2bb

2N
N
= 2 bb
2
R
R

(14.8)

such encounters with an impact parameter between b and b + b;


the mean quadratic velocity change is thus
!2
2Nbb 2Gm
8NG2 m2 b
2
v
=
;
(14.9)
R2
bv
R2 v2 b
integration this expression, we need to take into account that the
assumption of Borns approximation requires that
v . v

2Gm
.v
bv

b & bmin =

Gm
,
v2

(14.10)

and thus we obtain


!2
!2
Z
2Gm
2Gm
R
2
2
ln b|bmin 8N
ln ,
v =
v 8N
bv
bv
bmin
(14.11)
where
R
Rv2
ln ln
= ln
;
(14.12)
bmin
Gm
is the so-called Coulomb logarithm;
a typical velocity for the stars in a galaxy of mass M = Nm is,
according to the virial theorem,
v2

GMm
R

GNm
;
v2

(14.13)

from which we obtain


v2 8 ln

;
v2
N

(14.14)

this shows by what relative amount the stars velocity is changed


during one passage through the galaxy;

CHAPTER 14. JEANS EQUATIONS AND JEANS THEOREM 174


the Coulomb logarithm ln follows from
R
Rv2
ln = ln
= ln
ln N ,
bmin
Gm

(14.15)

i.e. the relative velocity change is approximated by


v2 8 ln N

;
v2
N

(14.16)

after ncross passages through the galaxy, the total relative velocity
change will approximately be
ncross

8 ln N
;
N

(14.17)

if this should be of order unity, the number of passages needs to


be
N
;
(14.18)
ncross
8 ln N
one passage takes approximately the time
tcross

R
,
v

(14.19)

i.e. a complete velocity change needs the relaxation time


R N
;
v 8 ln N

(14.20)

10 kpc
5 107 yr ,
200 km s1

(14.21)

trelax
in a galaxy, we have
tcross

and N 1011 , thus the relaxation time is


trelax 3 1016 yr ,

(14.22)

which is much more than the age of the Universe; this illustrates
that in many astrophysically relevant systems, the collision-less
Boltzmann equation can be used;
in a globular cluster, on the other hand, N 105 and tcross
105 yr, and thus
trelax 108 yr ,
(14.23)
which is short compared to the life time of the globular cluster; in
such cases, therefore, collisions do play a role;

CHAPTER 14. JEANS EQUATIONS AND JEANS THEOREM 175

14.2

The Jeans Equations

14.2.1

Moments of Boltzmanns equation

we thus begin again with Boltzmanns collision-less equation, in


which the right-hand side is set to zero,
df
=0
dt

(14.24)

consider f (~x,~v, t) as a function of position, velocity and time, and


replace the time derivative of the velocity according to Newtons
second law
F~
~
~v = =
(14.25)
m
to obtain
f
~ f
~ f = 0 ;
(14.26)
+ ~v
t
~v
as several times in the course of this lecture, we now form moments of equation (14.26) by integrating over velocity space,
Z
Z
Z

3
3
~ f
~ d3 v f = 0 ;
(14.27)
d v f + d v~v
t
~v
the last term here leads to boundary terms which vanish under the
assumption that there are no infinitely fast point masses,
f (~x,~v, t) 0

for |~v| ;

(14.28)

it is a divergence in velocity space to which Gau theorem can


be applied;
also, the gradient can be pulled out of the integral in the second
term, and this yields
Z
n ~
+ d3 v f~v = 0 ;
(14.29)
t
the mean velocity is defined as
1
h~vi =
n

d3 v, f~v ,

(14.30)

and so we find the continuitiy equation for the point masses in the
form
n ~
+ (nh~vi) = 0 ,
(14.31)
t
as expected;

CHAPTER 14. JEANS EQUATIONS AND JEANS THEOREM 176


the second moment of Boltzmanns equation is taken by multiplying the equation with ~v prior to integration over velocity space; in
this way, we obtain, written in components
Z
Z
Z

f
3 f i j
3
j
d v fv + d v iv v i
d3 v i v j = 0 ; (14.32)
t
x
x
v
partial integration of the third term yields
Z
Z
v j
3
3 f j
d v i v = d v f i = nij ;
v
v

(14.33)

this allows us to write (14.32) as


(nhv j i) (nhvi v j i)

+
n
=0,
+
t
xi
x j

(14.34)

where

Z
1
hv v i
d3 v f vi v j
n
is the correlation matrix of the velocity components;
i j

(14.35)

we now multiply the continuity equation (14.31) with v j ,


vj

(nhvi i)
n
+ vj
=0,
t
xi

(14.36)

subtract it from (14.34) and use the continuity equation to obtain


n

(nhvi i) (nhvi v j i)

hv j i
hv j i
+
= n j ;
i
i
t
x
x
x

(14.37)

the velocity-correlation matrix, hvi v j i, can be re-written to read


hvi v j i = h(vi hvi i)(v j hv j i)i + hvi ihv j i
(2 )i j + hvi ihv j i ,

(14.38)

with which we can cast (14.37) into the form


hv j i
n
hvi i
hvi ihv j i i nhv j i i
n
t
x
x

 n

 2 ij
2 ij
i
j
i
j
+ ( ) + hv ihv i
+
n
(
)
+
hv
ihv
i
xi
xi

= n j
(14.39)
x
which we can reduce to
n


hv j i
hv j i


+ nhvi i i = n j i (2 )i j n ;
t
x
x
x

(14.40)

CHAPTER 14. JEANS EQUATIONS AND JEANS THEOREM 177


for convenience, we now abbreviate
vi hvi i

(14.41)

because only averaged velocities and no velocities of individual


particles remain, and we obtain the two equations
n ~
+ (n~v) = 0 ,
t
i

 2 ij 
vi
j v
=

+v
( ) n ;
t
x j
xi x j

(14.42)

these are the Jeans equations which were derived for the first time
by Maxwell, but first applied to stellar-dynamical problems by
Sir James Jeans; obviously, the second equation corresponds to
Eulers equation in ideal hydrodynamics, where the divergence of
the tensor (2 )i j n takes the role of the pressure gradient,

~
P
1 P

= ij j j (2 )i j n ;

x
x

14.2.2

(14.43)

Jeans equations in cylindrical and spherical coordinates

it is useful for many applications to write the distribution function f as a function not of cartesian, but of curvilinear coordinates, such as cylindrical or spherical coordinates; for instance,
in cylindrical coordinates, we first have
(14.44)
x = r cos , y = r sin ,
x = r cos r sin , y = r sin + r cos
and thus in the plane perpendicular to the z axis
!
!

sin

cos
e
~v = r
+ r
= r~er + r~
cos
sin

(14.45)

as well as
x = r cos 2r sin r sin r 2 cos ,
y = r sin + 2r cos + r cos r 2 cos , (14.46)
and thus
e ;
~a = (r r 2 )~er + (2r + r)~

(14.47)

the gradient in cylindrical coordinates is


~ = ~er + ~e + ~ez ,

r
r
z

(14.48)

CHAPTER 14. JEANS EQUATIONS AND JEANS THEOREM 178


and we obtain
vr = r ,
ar =
vr =
a =
v =

v = r , vz = z ,

r r 2 =

r
2

v
2

+ r =
+
,
r
r
r
1

2r + r =
r
1 vr v

r + r =

, az = z = vz =
;
r
r
z

this implies the collision-less Boltzmann equation in cylindrical


coordinates,
df
f
f v f
f
=
+ vr
+
+ vz
dt
t
r
r
z
2

!
v f
vr v 1 f
f

r
r vr
r
r v z z
= 0;
(14.49)
in the same way, we can transform Boltzmanns equation to
spherical coordinates;
from Boltzmanns equation in spherical coordinates, we find after
integration over vr and under the practically important assumption
hv i = 0 = hv i

(14.50)

the equation
i
d(n2r ) n h 2
d
+ 2r (2 + 2 ) = n
,
dr
r
dr
where 2r,, are the velocity dispersions
Z
1
2
r,,
d3 v v2r,, ;
n

14.2.3

(14.51)

(14.52)

Application: the mass of a galaxy

as an application of the Jeans equations, we consider a galaxy in


spherical coordinates whose polar and azimuthal velocity dispersions are assumed to be equal,
2 = 2 ;

(14.53)

we introduce the anisotropy parameter


2
2
1 2 =1 2 ,
r
r
which is typically 0;

(14.54)

CHAPTER 14. JEANS EQUATIONS AND JEANS THEOREM 179


the spherically-symmetric Jeans equation then reads
d(n2r ) 2n 2
d
+
r = n
;
dr
r
dr

(14.55)

for a spherically-symmetric system, we can write


d GM(r)
,
=
dr
r2

(14.56)

and thus we find


"
#
d2r 2n 2
GM(r)
r 2 dn
2
= vc = r
+n
+

r2
n
dr
dr
r r
#
"
d ln 2r
2 d ln n
+
+ 2 ;
(14.57)
= r
d ln r
d ln r
here, vc is the orbital velocity on a circular orbit with radius r
around the centre of the galaxy; given an assumption for , such
as = 0, this equation allows us to determine the mass of a
galaxy, for instance if the surface-brightness profile is used as
a measure for d ln n/d ln r and the profile of the radial velocity
dispersion is measurable;

14.3

The Virial Equations

14.3.1

The tensor of potential energy

we return to the Jeans equation in the form (14.34)

(nhv j i) (nhvi v j i)
+
+n
=0;
i
t
x
x j

(14.58)

multiplication with the particle mass m and the spatial coordinates


xk , followed by integration over d3 x yields
Z
Z
Z
i j
j

3
k (hv v i)
3
k (hv i)
= d xx
d3 x xk
;
d xx
i
t
x
x j
(14.59)
the second term on the right-hand side is Chandrasekhars tensor
of the potential energy,
Z

i
Wj
d3 x xi j ,
(14.60)
x
whose trace is the systems potential energy;

CHAPTER 14. JEANS EQUATIONS AND JEANS THEOREM 180


the properties of W ij become clearer if we insert the potential explicitly,
Z
(~x0 )
= G d3 x0
,
(14.61)
|~x ~x0 |
and use the gradient
Z
j
0j
x0 )

3 0 (x x )(~
=
G
d
x
;
(14.62)
x j
|~x ~x0 |3
then, Chandrasekhars tensor becomes
Z
Z
xi (x j x0 j )
i
3
;
W j = G d x d3 x0 (~x)(~x0 )
|~x ~x0 |3

(14.63)

we now swap ~x and ~x0 and change the order of integrations, obtain
Z
Z
x0i (x j x0 j )
i
3
W j = +G d x d3 x0 (~x)(~x0 )
;
(14.64)
|~x ~x0 |3
and add this to the previous expression (14.63) to find
Z
Z
G
(xi x0i )(x j x0 j )
i
3
Wj =
d x d3 x0 (~x)(~x0 )
; (14.65)
2
|~x ~x0 |3
first, this shows that W ij is manifestly symmetric, W ij = Wij , and
its trace is the potential energy,
Z
Z
Z
x)(~x0 ) 1
G
3
3 0 (~
i
d x d x
=
d3 x (~x)(~x) ,
Wi =
0
2
2
|~x ~x |
(14.66)
as claimed;

14.3.2

The tensor virial theorem

now we return to the first term on the right-hand side of the spatial
integral (14.58),
Z
Z
i j
k
i j
3
k (hv v i)
3 (x hv v i)
d xx
=
d
x
xi
xi
Z
xk

d3 x hvi v j i i ; (14.67)
x
the first term on the right-hand side is a divergence and vanishes
upon integration over a closed system; the second is the tensor of
the kinetic energy, multiplied by 2,
Z
1
i
Kj
d3 x hvi v j i ;
(14.68)
2

CHAPTER 14. JEANS EQUATIONS AND JEANS THEOREM 181


we insert the definition (14.38) of (2 )ij into this tensor and obtain
1
1
K ij = T ij + ij ,
2
2
where T ij and ij are the tensors
Z
Z
i
3
i
i
Tj
d x v v j , j
d3 x (2 )ij ,

(14.69)

(14.70)

where the mean velocity hvi i was written as vi again;


obviously, the tensor T ij corresponds to the stress-energy tensor in
ideal hydrodynamics up to the pressure term, while ij describes
the momentum transport by unordered motion and thus a form of
anisotropic pressure;
on the left-hand side of the spatial integral (14.58), the term
Z
(hv j i)
(14.71)
d3 x xk
t
remains; we use its symmetry to write it as
"
#
Z
j
k
1
3
k (hv i)
j (hv i)
d x x
+x
2
t
t

(14.72)

or

Z


1
d3 x xk v j + x j vk ;
(14.73)
2 t
the partial time derivative can again be replaced by a total time
~ vanishes when
derivative because the convective derivative ~v
applied to the volume integral, and this finally yields
Z


1d
d3 x xi v j + x j vi = T ij + ij + W ij ,
(14.74)
2 dt

where the symmetry of the three tensors T ij , ij and W ij was used


again;
from our earlier considerations on the tensor virial theorem, we
know that
Z

 d2 I ij
d
i
i
3
d x x v j + x jv = 2 ,
(14.75)
dt
dt
where I ij is the tensor of second moments of the mass distribution,
Z
i
Ij
d3 x xi x j ;
(14.76)
thus, we obtain the tensor virial theorem for collision-less systems,


d2 I ij
i
i
i
=
2
T
+

+
W
(14.77)
j
j
j ;
dt2

CHAPTER 14. JEANS EQUATIONS AND JEANS THEOREM 182


taking the trace of this equation leads us back to the ordinary
(scalar) virial theorem, if the mass distribution is static,
d2 Iii
=0
dt2

T ii + ii + Wii = 0 ;

(14.78)

now, the trace of the sum


T ii

ii

2Kii

d3 x v2

(14.79)

is twice the kinetic energy, and Wii is the total potential energy, as
we saw before; thus
2K = W ,
(14.80)
which is the scalar virial theorem;

14.4

The Jeans Theorem

an integral of motion in a gravitational field is a quantity f


which satisfies
d f (~x,~v)
=0,
(14.81)
dt
where ~x(t) and ~v(t) are characterising arbitrary orbits; by means
~ this can be cast into the form
of the equation of motion, ~v = ,
d f (~x,~v) ~
f
~ f
~ f = 0 ;
= ~v
= ~x f + ~v
dt
~v
~v

(14.82)

in comparison to the collision-less Boltzmann equation, we see


that f is an integral of motion if and only if it is a stationary
solution of Boltzmanns equation,
f
=0;
t

(14.83)

this leads us to Jeans theorem: Any stationary solution of the


collision-less Boltzmann equation depends on the phase-space
coordinates only through integrals of motion, and conversely any
function depending only on integrals of motion is a stationary solution of the collision-less Boltzmann equation;
the proof of the first statement has already been given; as to the
second statement, let Ii , 1 i n be n integrals of motion and
f (I1 , I2 , . . . , In ) an arbitrary function thereof; then,
df
f dIi
=
=0,
dt
Ii dt
and f solves the collision-less Boltzmann equation;

(14.84)

CHAPTER 14. JEANS EQUATIONS AND JEANS THEOREM 183


an orbit in a gravitational potential always has six integrals of motion; namely, let the orbit be specified by ~x(t) and ~v(t), then it can
be traced back to the initial point ~x0 , ~v0 by means of the equations
of motion; these six numbers can be considered as integrals of
motion;
one distinguishes isolating and non-isolating integrals of motion;
isolating integrals such as the energy E or the angular momentum
~L constrain the orbits; an orbit passing through the point ~x0 , ~v0
in phase space is part of a subspace S n of phase space defined
by the 6 n conditions I1 = const., . . . , I6n = const.; an integral
I(~x,~v) is isolating for this orbit if an n-dimensional subspace S n
exists in which no point comes arbitrarily close to the hypersurface I(~x,~v) = I(~x0 ,~v0 );
isolating integrals are extraordinarily important, non-isolating orbits have no practical importance;

Chapter 15
Equilibrium, Stability and
Disks
15.1

The Isothermal Sphere

15.1.1

Phase-space distribution function

further reading:
Binney,
Tremaine, Galactic Dynamics,
sections 4.5, 4.7, 5.1 and 5.3

spherical systems which are independent of time have orbits


with at least the four integrals of motion energy E and angular momentum ~L; the Jeans theorem then tells us that any (nonnegative) function f (E, ~L) is a stationary solution of the collisionless Boltzmann equation and may thus represent a stable, selfgravitating system;
the gravitational potential caused by a system with phase-space
distribution function f is determined by Poissons equation,
Z
2
~
= 4G = 4Gm d3 v f ;
(15.1)
if the system is isotropic also in velocity space, it cannot depend
on the direction of angular momentum, which implies
f (E, ~L) = f (E, L) ;
using the Laplacian operator in spherical symmetry,
!
1
d
d
2
2
~ =
r
,

r2 dr
dr
the equation for the gravitational potential
!
!
Z
1 d 2 d
mv2
3
r
= 4Gm d v f
+ m, m|~x ~v|
r2 dr
dr
2

(15.2)

(15.3)

(15.4)

follows as the fundamental equation for self-gravitating spherical


systems in equilibrium;
184

CHAPTER 15. EQUILIBRIUM, STABILITY AND DISKS

185

it is often convenient to re-scale the Potential and the energy E


by subtracting a constant 0 ,
+ 0 ,

E
v2
+ 0 = ;
m
2

(15.5)

we now consider a simple example, in which f does not depend


on L,
!
n1
n1
v2 /2
E/2
f (E) =
e
=
exp
, (15.6)
(22 )3/2
(22 )3/2
2
with a normalising constant n1 ;
integration over all velocities yields
4n1 e/
(22 )3/2

4n1 e/
=
(22 )3/2

2 v2 /(22 )

dv v e

(22 )3/2
4

= n1 e/ n ,
2

(15.7)

where n is the number density of particles;

15.1.2

Isothermality

Poissons equation for this system reads


!
1 d 2 d
2
r
= 4Gnm = 4Gm1 e/ ,
2
r dr
dr

(15.8)

or, using

n
= 2 (ln n ln n1 ) ,
n1
we find an equation for the number density n,
!
1 d 2 dn
4G
r
= 2 n,
2
r dr
dr

= 2 ln

(15.9)

(15.10)

which can of course also be considered as an equation for the


density = nm;
in (ideal) hydrodynamics, we had derived the equation
!
kT r d ln gas d ln T
M(r) =
+
mG d ln r
d ln r

(15.11)

for a hydrostatic, spherical system; if this is isothermal, dT/dr =


0, we first have
Z
d ln gas
mG r 0 02
2
r
=
dr r gas (r0 )
(15.12)
dr
kT 0

CHAPTER 15. EQUILIBRIUM, STABILITY AND DISKS

186

if we consider the mass as only contributed by the gas (without


any dark matter); differentiation with respect to r and division by
r2 yields
!
1 d 2 d ln gas
4Gm
r
=
gas ;
(15.13)
2
r dr
dr
kT
this equation is identical with (15.10) which we have just derived
from the Jeans theorem, if we set
2

kT
;
m

(15.14)

thus, the corresponding stellar-dynamical model is called the


isothermal sphere;
the mean-squared velocity in the isothermal sphere is
 2
R
Z
v
dv v4 exp 2
2
1
2
3 2
 2
hv i =
d vv f = R
v
n
dv v2 exp 2
2


3
1 5/2
3
8
22
=  3/2 = 2 = 32 ;

2
1
4

(15.15)

22

any individual velocity component thus has the mean square


hv2x i = hv2y i = hv2z i = 2 ;

15.1.3

(15.16)

Singular and non-singular solutions

one solution of the equation (15.10) for the density of the isothermal sphere follows from the ansatz
n = Cr ;

(15.17)

on the left-hand side of the equation, it yields

,
r2

(15.18)

and thus the ansatz indeed is a solution if = 2 and C =


2 /(2G),
2
= mn =
;
(15.19)
2Gr2
this solution is called the singular isothermal sphere;
another solution which avoids the central singularity can be
found numerically; for doing so, we conveniently introduce the
dimension-less quantities
x

r
,
r0

,
0

(15.20)

CHAPTER 15. EQUILIBRIUM, STABILITY AND DISKS

187

where 0 is meant to be the finite central density; then, the equation for the scaled density y is
!
d 2 d ln y
4G
(15.21)
x
= 2 0 r02 yx2 ;
dx
dx

if we set the scale radius r0 to


s
r0

92
,
4G0

(15.22)

the equation simplifies to


!
d 2 d ln y
x
= 9yx2 ;
dx
dx

(15.23)

this equation can be integrated with the boundary conditions



dy
=0;
(15.24)
y(0) = 1 ,
dx 0
within x . 2, i.e. r . 2r0 , the numerical result can be approximated by
1
y(x)
;
(15.25)
(1 + x2 )3/2

15.2

Equilibrium and Relaxation

is there an equilibrium state of a self-gravitating system, which


corresponds to an entropy maximum? the entropy
Z
S
d3 xd3 p p ln p
(15.26)
phase space
is maximised if and only if p is the distribution function of the
isothermal sphere; however, the isothermal sphere has infinite
mass and energy and can thus not be an exact description of a
thermodynamical equilibrium state; this implies that there is no
thermodynamical equilibrium of a self-gravitating system, and
that self-gravitating systems cannot have stable final configurations, but at best long-lived transient states!
if we populate a narrow region in phase space with N stars, their
orbits will have slightly different initial conditions; as time proceeds, they will progressively evolve away from each other and
thus occupy a growing part of phase space; this phase mixing

CHAPTER 15. EQUILIBRIUM, STABILITY AND DISKS

188

causes the averaged phase-space distribution f to decrease, because the averaged phase-space density is progressively diluted;
thus, the macroscopic entropy
Z
S d3 xd3 v f ln f
(15.27)
does indeed increase;
this process of phase mixing is in fact hardly different from the
thermodynamical trend to equilibrium; there, too, the increase of
entropy is caused by macroscopically averaging over processes
which are otherwise reversible;
if the potential is changed while the particles are moving through
it, energy can be transported from particles to others; if, for example, the system contracts while a star approaches its centre, the
potential deepens and the star looses energy; other stars can gain
considerable amounts of energy; this process is called violent relaxation (Lynden-Bell);

15.3

Stability

15.3.1

Linear analysis and the Jeans swindle

as before in hydrodynamics, we consider an equilibrium solution f0 , 0 of the coupled system of the collision-less Boltzmann
equation and Poissons equation,
f
~ f
~ f = 0 ,
+ ~v
t
Z~v
~ 2 = 4Gm d3 v f ;

(15.28)

in equilibrium, f /t = 0;
then, we perturb f0 , 0 by small amounts f , and linearise the
equations in f , :
f
~ f
~ 0 f
~ f0 = 0 ,
+ ~v
t
~v
Z ~v
~ 2 = 4Gm d3 v f ;

(15.29)

as an equilibrium solution, we adopt an infinitely extended, homogeneous distribution f0 , which implies a density 0 and a potential 0 given by
~ 2 0 = 4G0 ;

(15.30)

CHAPTER 15. EQUILIBRIUM, STABILITY AND DISKS

189

due to the infinite extend and the homogeneity, we must have


~ 0=0,

(15.31)

which is possible only if 0 = 0; we invoke the Jeans swindle


to set 0 = 0; this is practically permissible if we study perturbations which are small in scale compared to possible scales in
0 ;
with the Jeans swindle, the linearised equations read
f
~ f
~ f0 = 0 ,
+ ~v
t
Z ~v
~ 2 = 4Gm d3 v f ;

(15.32)

again, we decompose the solution into plane waves,


~

f = fv (~v)ei(k~xt) ,

= v ei(k~xt)

(15.33)

and obtain as a condition for f to solve the equations


f0
=0
i fv + i~v ~k fv iv~k
Z ~v
k2 v = 4Gm d3 v fv ;

(15.34)

the first equation yields


fv = v~k

1
f0
~v ~k ~v

(15.35)

which, when inserted into the second, yields


k v = 4Gm

dv
3

v~k

f0
~v

~k ~v

(15.36)

since v does not depend on ~v, we find


4Gm
1+
k2

d3 v

~k

f0
~v

~k ~v

=0

(15.37)

this correponds exactly to the longitudinal dielectricity l from


plasma physics,
4e2
l = 1 2
k

Z
3

d p

~k

f0
~p

~k ~v

(15.38)

and thus leads to Landau damping, exactly as in plasma physics;

CHAPTER 15. EQUILIBRIUM, STABILITY AND DISKS

15.3.2

190

Jeans length and Jeans mass

the boundary between stable an unstable solutions is defined by


= 0; if we suppose a thermal system with a Maxwellian velocity distribution,
n0
2
2
ev /(2 ) ,
f0 (~v) =
(15.39)
2
3/2
(2 )
we have
~v
f0
(15.40)
= f0 (~v) 2 ;

~v
if ~k is chosen parallel to the positive x axis, the condition
Z
v2x /(22 )
4Gmn0
3 kv x e
1 2 2
dv
(15.41)
k (22 )3/2
kv x
follows; for = 0, the integral is
Z
2
2
d3 v evx /(2 ) = (22 )3/2 ,
(15.42)
and we find
4G0
4G0
1 2 2 = 0 k2 ( = 0) kJ2 =
; (15.43)
k
2
instability sets in for smaller k or wave lengths larger than J =
2/kJ ; the quantity

2
2

= p
=
J
(15.44)
kJ
G0
4G0
is called the Jeans length;
the Jeans length defines the volume
!3

3
(15.45)
J =
G0
and thus the mass
!1/3
!1/3
GM2J
MJ
3
MJ 0 J

J
=
0
2
GM

J
(15.46)
;
2
according to the virial theorem,
GM
GM
2
R 2 ;
(15.47)
R

the radius of the system is thus comparable to the Jeans length;


this means that the assumption of homogeneity on the scale of
the Jeans length cannot be satisfied and that the nature of the instability needs to be studied for each system in detail once its
geometry is specified; nonetheless, the Jeans length defines an order of magnitude estimate for the boundary between stability and
instability;

CHAPTER 15. EQUILIBRIUM, STABILITY AND DISKS

191

15.4

The rigidly rotating disk

15.4.1

Equations for the two-dimensional system

as an example for a rotating system with flat geometry, we consider an infinitely thin disk (with thickness zero) which is rigidly
~ = ~ez ; the
rotating around the z axis with an angular velocity
disk thus fills the x-y plane and have a surface-mass density 0 ;
we consider perturbations in the plane of the disk and neglect
warps or twists; furthermore, we transform into a co-rotating coordinate frame and study the disk in the (simpler) fluid approximation;
then, the continuity equation, Eulers and Poissons equations
read
~
+ (~v) = 0 ,
t
~
~v
~ v = P
~ 2
~ ~v +
~ 2~r ,
+ (~v )~
t

~ 2 = 4GD (z) ;

(15.48)
here, we had to take into account in Eulers equation that Coriolis
and centrifugal forces occur in the co-rotating coordinate frame;
the physical quantities occuring here are two-dimensional,
~v(x, y, t), (x, y, t) and so on, and for the pressure we assume a
barotropic equation-of-state,
P(x, y, t) = P[(x, y, t)] ;

(15.49)

the unperturbed quantities are obviously ~v = 0 and = 0 as


well as P0 = P(0 ); this trivially satisfies the continuity equation,
Eulers equation reads
~ 0 = 2~r ,

(15.50)

~ 2 0 = 4G0 D (z) ;

(15.51)

and Poissons equation is

~ 0
since no direction can be preferred on a homogeneous disk,
must point along the z axis, which contradicts Eulers equation;
thus, there is no gravitational force yet to balance the centrifugal force; therefore, we assume that the disk is embedded into a
surrounding gravitational field which compensates the centrifugal
force, such as the halo of a galaxy;

CHAPTER 15. EQUILIBRIUM, STABILITY AND DISKS

15.4.2

192

Analysis of perturbations

we now perturb the disk by small amounts , ~v and so on and


linearise the equations in the perturbations; this implies

~ ~v = 0 ,
+ 0
t
c2 ~
~v
~
~ ~v ,
= s

2
t
0
~ 2 = 4GD (z) ,
(15.52)

where the sound velocity was introduced as


!
dP()
2
cs =
;
d 0

(15.53)

as usual, we decompose the perturbations into plane waves,


~

~v = ~vA ei(k~xt) ,
~

= A ei(k~xt) ,
~

= A ei(k~xt)

(15.54)

valid at z = 0, and turn the x axis into the direction of ~k;


we first consider Poissons equation; for z = 0, we have
= A ei(kxt) ,

(15.55)

=0

(15.56)

while

must hold otherwise; this is achieved by


= A ei(kxt)|kz| ,

(15.57)

where k = k x can have either sign;


we now integrate Poissons equation along the z direction from
to and then take the limit 0; due to the continuity of
2
x2

and

2
y2

(15.58)

~ 2 , but what remains


at z = 0, these two terms disappear from
is
Z

2
!
= 4G ;
lim
dz
(15.59)
= lim
0
0 z
z2
however, we also have


= 2|k|A ei(kxt) ,
lim
0 z

(15.60)

CHAPTER 15. EQUILIBRIUM, STABILITY AND DISKS

193

and so A is related with A through


2|k|A = 4GA ;

(15.61)

thus, we find
=

2GA i(kxt)|kz|
e
;
|k|

(15.62)

putting these equations into the continuity and Eulers equations,


we find
i0 kvAx = iA ,
c2
2Gk
ivAx = i s kA + i
A + 2vAy ,
0
|k|
(15.63)
ivAy = 2vAx ;
these are three linear equations for the three variables A , vAx
and vAy ; writing them in the form

0
 i  ik0
A

2
ik cs 2G i 2 vAx = 0
(15.64)
0
|k|


v
Ay
0
2 i
makes it immediately obvious that non-trivial solutions exist if
and only if the determinant of the matrix vanishes, hence
!
#
"
c2s 2G
2
2

i( 4 ) + ik0 ik
(i) = 0 ;
(15.65)
0
|k|
this means that either = 0 or
2 = 42 + k2 c2s 2G|k|0 ;

(15.66)

this dispersion relation describes the modes of the perturbed,


rigidly rotating disk; the modes are stable for 2 0 and unstable
for 2 < 0;

15.4.3

Toomres criterion

if = 0, which is boring for a rotating disk, the disk is unstable


if
2G0
|k| < kJ
,
(15.67)
c2s
where kJ is the Jeans wave number for the disk; if the sound speed
can be arbitrarily low, cs 0, perturbations are unstable for arbitrarily large k; their growth rate is then
= 2G0 |k|

(15.68)

and the growth is exponential, et ; obviously, small perturbations with |k| , 0, grow particularly violently, i.e. the
cold, non-rotating disk fragments violently on small scales;

CHAPTER 15. EQUILIBRIUM, STABILITY AND DISKS

194

this is not suppressed by rotation either; for cs 0, 2 < 0 for


22
|k| >
,
G0

(15.69)

i.e. even then the instability sets in on the smallest scales;


pressure and rotation are obviously not able individually to stabilise the disk, since for = 0, the disk is unstable for perturbations with small, and for P = 0 = cs for large wave lengths;
however, pressure and rotation can be stabilising if they act together, since then the dispersion relation has a minimum where
i !
h 2 2
|k| cs 2G0 |k| + 42 2 = 0 ,
k

(15.70)

which yields
2|k|c2s = 2G0

|k| =

G0 kJ
=
;
c2s
2

(15.71)

obviously, the disk is stable if and only if 2 0 at this wave


number because it is then positive for all wave numbers; thus, the
condition for global stability is
!
G0
G0 (G0 )2
2
2
> 0 (15.72)
4 2G0 2 +
= 4
cs
c2s
c2s
or

cs
> 1.57 ;
(15.73)
G0 2
this is Toomres criterion, which can also be applied to collisionless systems (recall that we had adopted the fluid approximation!); then,
cs
& 1.68
(15.74)
G0
is the condition for stability;

Chapter 16
Dynamical Friction,
Fokker-Planck Approximation
16.1

Dynamical Friction

16.1.1

Deflection of point masses

further reading:
Binney,
Tremaine, Galactic Dynamics,
sections 7.1 and 8.3

an interesting effect occurs if a mass M moves through an environment of masses m which are homogeneously distributed
around the mass M; although the motion of the masses can be
considered collision-less, a deceleration occurs which is called
dynamcial friction;
let ~vm and ~v M be the velocities of one of the masses m and of the
mass M, respectively; ~xm and ~x M are their locations; further,
~r ~xm ~x M

(16.1)

is the distance vector from M to m, and


~v ~r = ~vm ~v M

(16.2)

is the relative velocity of m with respect to M; the system of two


point masses obeys the equation of motion around a fixed force
centre of a single body with the reduced mass,
 mM 
GMm

~r = 2 ~er 2 ~er ,
(16.3)
m+M
r
r
where ~er is the unit vector in radial direction away from M;
obviously, the change of ~v equals the difference of the changes in
~vm and ~v M ,
~v = ~vm ~v M ,
195

(16.4)

CHAPTER 16. DYNAMICAL FRICTION, FOKKER-PLANCK APPROXIMATION196


and, furthermore, the velocity of the centre of mass is unchanged,
~ m~xm + M~x M
X
and thus

~ = 0
X

~ = m~vm + M~v M ,
X

m~vm + M~v M = 0 ;

(16.5)

(16.6)

consequently then, ~vm = ~v + ~v M , and thus


m(~v + ~v M ) = M~v M

~v M =

m
~v ;
m+M

(16.7)

we shall now determine ~v;


the fictitous particle with the reduced mass, Mm/(M + m), now
describes a hyperbolic orbit around the (resting) centre of force;
from the Kepler problem of celestial mechanics, we know that the
complete scattering angle is given by
sin

1
= ,
2 

(16.8)

where  is the orbits eccentricity;


this implies that the cosine of (half) the scattering angle is
r
r

1
1 2
cos = 1 sin2 = 1 2 =
 1,
(16.9)
2
2


and thus
tan

1
=
;
2
2 1

(16.10)

generally, the treatment of the Kepler problem shows that the eccentricity is related to energy E and angular momentum ~L through
s
2L2 E
 = 1+ 2 ;
(16.11)

if the impact parameter is b, the angular momentum is


mM
bv bv
m+M

(16.12)

mM v2 2
E=
v
m+M 2
2

(16.13)

L=
and the energy is

because we need to insert the reduced mass

mM
m+M

(16.14)

CHAPTER 16. DYNAMICAL FRICTION, FOKKER-PLANCK APPROXIMATION197


for the mass of the fictitous particle whose motion we study; since
GMm = G(m + M) ,
the eccentricity is
s
s
2
2
2v (bv)
b2 v4
=
1
+
,
 = 1+
2[G(M + m)]2
G2 (M + m)2

(16.15)

(16.16)

and thus the tangent of (half) the scattering angle is


tan

1
= q
2
b2 v4

G(M + m)
;
bv2

(16.17)

G2 (M+m)2

16.1.2

Velocity changes

because of energy conservation, the velocity of the fictitous mass


is v also for t , and because of the deflection by the angle ,
the velocity components parallel and perpendicular to the asymptotic incoming direction are
2 tan /2
1 + tan2 /2
2G(M + m)
1
= v
h G(M+m) i2
2
bv
1 + bv2

v = v sin = v

(16.18)

b2 v4
G(M + m)
bv
b2 v4 + G2 (M + m)2

1
#2
"

bv3
bv2

= 2
1+
;

G(M + m)
G(M + m)

= 2

the velocity change of the mass M perpendicular to its initial direction of motion is thus
m
v
(16.19)
v M =
M+m

"
#2 1

2bmv3
bv2

=
1
+
;

G(M + m)2
G(M + m)
in parallel direction, we have
vk = v cos

vk = v(cos 1) = v(1 cos ) (16.20)

or
vk

!
1 tan2 /2
2 tan2 /2
=
v
= v 1
1 + tan2 /2
1 + tan2 /2
2v
2v
=
=
(16.21)
h 2 i2 ;
2
1 + tan /2 1 + bv
G(M+m)

CHAPTER 16. DYNAMICAL FRICTION, FOKKER-PLANCK APPROXIMATION198


the velocity change in parallel direction is thus
v Mk =

m
2mv
1
vk =
h 2 i2 ;
M+m
M + m 1 + bv

(16.22)

G(M+m)

if the mass M is moving through a homogeneous sea of masses


m, all perpendicular deflections must cancel, while the parallel
velocity changes must add up; therefore, the mass M will experience a steady deceleration;

16.1.3

Chandrasekhars formula

let f (~vm ) be the phase-space density of the stars with mass m


which constitute the sea of masses; then, the rate at which the
masse M encounters collisions with stars with an impact parameter between b and b + db is
2bdb v f (~vm )d3 vm ;

(16.23)

these collisions change the velocity of M by


d~v M
dt

= ~v f (~vm )d3 vm
bmax

(16.24)

2mv
db 2b
m+M

1
#2
"

bv2

1+

G(M + m)

where bmax is the maximum possible impact parameter, which


could be defined by the physical size of the cloud of stars m;
the integral
Z

bmax

db
0

b
1+

bv2

i2 =

G(M+m)

bmax

db
0

1
=
2a

Z
1

b
1 + ab2

1+ab2max

d
, (16.25)

where 1 + ab2 was set; thus,


Z

bmax

db
0

where

b
1+

i2
bv2
G(M+m)

"
#2
1 G(M + m)
=
(16.26)
2
v2
"
#
b2max v4
,
ln 1 + 2
G (M + m)2

v2
a
G(M + m)
"

#2
(16.27)

CHAPTER 16. DYNAMICAL FRICTION, FOKKER-PLANCK APPROXIMATION199


was used; let further

bmax v2
,
G(M + m)

(16.28)

then the rate of change for the velocity of M by encounters with


stars of mass m and velocity ~vm is
"
#2
d~v M
2mv G(M + m)
3
ln(1 + 2 )
= ~v f (~vm )d vm
2
dt
m+M
v
2
2
= 2G ln(1 + )m(M + m)
~vm ~v M
f (~vm )d3 vm
,
(16.29)
|~vm ~v M |3
where we have inserted ~v = ~vm ~v M ;
the quantity is typically  1, thus
ln(1 + 2 ) ln 2 = 2 ln ;

(16.30)

hence, we replace from now on ln(1 + 2 ) 2 ln ; typical values


for this so-called Coulomb logarithm are
5 . ln . 20 ;

(16.31)

for stars which are isotropically distributed in velocity space, the


integral over velocity space is determined by
Z
Z
~vm ~v M
~v M vM
2
dvm v2m f (vm ) , (16.32)
dvm vm f (vm )
= 3
3
|~vm ~v M |
vM 0
0
in perfect analogy to the gravitational field of a collection of mass
points with mean mass density ;
together, this implies Chandrasekhars formula for dynamical
friction,
Z
~v M vM
d~v M
2
2
dvm v2m f (vm ) ; (16.33)
= 16 ln G m(M + m) 3
dt
vM 0
if v M is small compared to the typical velocity of the stars m, the
remaining integral can be approximated by
Z vM
v3M
2
dvm vm f (vm )
f (0) ;
(16.34)
3
0
then,
d~v M
162G2
=
ln m(M + m) f (0)~v M ;
dt
3

(16.35)

the friction is then proportional to ~v M like for Stokes-type friction;

CHAPTER 16. DYNAMICAL FRICTION, FOKKER-PLANCK APPROXIMATION200


for sufficiently large v M , the integral converges to a constant,
which implies that the friction force becomes proportional to v2
M;
for a Maxwellian velocity distribution,
n0
2
2
ev /(2 ) ,
f (vm ) =
2
3/2
(2 )

(16.36)

the friction force becomes


"
#
d~v M
4G2 ln (M + m)0
2X X2
~v M , (16.37)
=
erf(X) e
dt
v3M

where X v M /( 2) scales the velocity v M ; here, we have set


0 mn0 ;
if M  m, (M + m) M, and the friction only depends on the
density 0 of the scatteres stars, but not on their mass any more,
"
#
d~v M
2X X2
4G2 ln 0 M
~v M ;
erf(X) e
=
(16.38)
dt
v3M

similarly, in this case the friction force os proportional to M 2 because the deceleration is proportional to M;

16.2

Fokker-Planck Approximation

16.2.1

The master equation

so far, we have considered the collision-less Boltzmann equation,


df
=0;
(16.39)
dt
in presence of collisions, it has to be augmented by collision terms
on the right-hand side,
df
= C[ f ] ,
(16.40)
dt
which can be written as a functional of the phase-space distribution function f ; explicitly, this equation reads
f ~
~ f = C[ f ] ;
+ ~x f
t
~v

(16.41)

~ in phase space
collisions transport particles from one position w
~ + ~
t another position w
w; within a time interval t, this may
happen with a probability
(~
w, ~
w)d6 ~
wt ;

(16.42)

like in hydrodynamics, we distinguish between long-ranged,


smooth forces short-ranged collisions;

CHAPTER 16. DYNAMICAL FRICTION, FOKKER-PLANCK APPROXIMATION201


~ are
the losses at w

Z
f (~
w)
w) d6 ~
w (~
w, ~
w) ,
= f (~
t
while the gain is

Z
f (~
w)
w ~
w) d6 ~
w (~
w ~
w, ~
w) ;
= f (~
t +
their sum is the total change


f (~
w) f (~
w)
f (~
w)
=
+
= C[ f ] ,
t
t +
t

(16.43)

(16.44)

(16.45)

which yields the so-called master equation


df
=
Zdt

(16.46)



d6 ~
w (~
w ~
w, ~
w) f (~
w ~
w) (~
w, ~
w) f (~
w) ;

16.2.2

The Fokker-Planck equation

most collisions will change the velocity only by a small amount;


we had seen in the beginning that
v

2Gm Gm 2v
bmin
= 2
= 2v
,
bv
v b
b

(16.47)

and thus

v bmin

;
(16.48)
v
b
the relative velocity change thus decreases like b1 while the number of collisions increases with b proportional to b2 ; thus most
collisions cause only small velocity changes;
we can use that to simplify the master equation; for small v, ~
w
is also small, and the first term under the integral in the master
equation can be expanded into a Taylor series,
(~
w ~
w, ~
w) f (~
w ~
w) (~
w, ~
w) f (~
w)


(~
w, ~
w) f (~
w) wi +
i
w

1 2 
(16.49)
(~
w, ~
w) f (~
w) wi w j ,
i
j
2 w w
if we stop after the second order; now, we can integrate over ~
w,
which yields the scattering term
"
#
Z

6
i
C[ f ] = i f (~
w) d w w (~
w, ~
w)
(16.50)
w
"
#
Z
1 2
6
i
j
+
f (~
w) d w w w (~
w, ~
w) ;
2 wi w j

CHAPTER 16. DYNAMICAL FRICTION, FOKKER-PLANCK APPROXIMATION202


we introduce the diffusion coefficients in phase space,
Z
i
D(w )
d6 w wi (~
w, ~
w)
and
D(w w )
i

d6 w wi w j (~
w, ~
w)

(16.51)

(16.52)

which enable us to bring the master equation into the FokkerPlanck form
f ~
~ f
+ ~x f
t
~v

i
h
i
f
(~
w
)D(w
)
(16.53)
wi
h
i
2

f (~
w)D(wi w j ) ;
i
j
w w
=

the great advantage of the Fokker-Planck approach is that the diffusion approximation in phase space depends only on the local
~ of a test particle, such that the integrophase space coordinates w
differential master equation turns into a pure differential equation;
the diffusion coefficients can now further be approximated by several simplifying assumptions;

Você também pode gostar