Radiative Gas Dynamics - Barbara Ryden
Radiative Gas Dynamics - Barbara Ryden
Radiative Gas Dynamics - Barbara Ryden
Due dates for problem sets = January 20, February 3, February 17, March 3
Astronomy 825:
Radiative Gas Dynamics
Winter 2011
Barbara Ryden
January 3, 2011
ii
Contents
6 Ionization Fronts 55
7 Basic Turbulence 67
8 Spherical Accretion 79
iii
iv CONTENTS
Notes on the Notes
These lecture notes are the result of my having taught Astronomy 825 several
times, starting in 1993, when my knowledge of radiative gas dynamics was
nearly non-existent. The most valuable resource I found while cramming to
teach the course was the two-volume Physics of Astrophysics set by Frank
Shu. Volume I of Shu’s work is titled Radiation and Volume II is titled Gas
Dynamics, and between them they cover more than you might want to know
about radiative gas dynamics. More recent works that I have consulted while
updating these notes are An Introduction to Astrophysical Fluid Dynamics
by Michael Thompson and Principles of Astrophysical Fluid Dynamics by C.
J. Clarke and R. F. Carswell.
In a way, these (not-a-textbook) notes are a companion to Rick Pogge’s
(not-a-textbook) notes for Astronomy 871, Physics of the Interstellar Medium.
I have attempted to make notation and units consistent between the two sets
of notes. Thus, I adopt cgs units, supplemented by angstroms (Å ), electron
volts (eV), and parsecs (pc).
v
vi NOTES ON THE NOTES
λ≪L (1.1)
where λ is the mean free path of the gas particles, and L is the characteristic
size of the system. The mean free path in a gas of neutral particles is
1
λ= (1.2)
nσ
where n is the number density of particles, and σ is the cross section for col-
lisions. A typical cross section for atoms or small molecules is σ ∼ 10−15 cm2 ,
or about 1 gigabarn. In air at room temperature, n ∼ 1019 cm−3 , and
hence λ ∼ 10−4 cm ∼ 1 µm. Thus, a volume of air that is larger than
several microns on a side can be treated as a continuous fluid. In the in-
terstellar medium (ISM), particle number densities are much lower than in
the Earth’s atmosphere. In a molecular cloud, n ∼ 1000 cm−3 , and hence
λ ∼ 1012 cm ∼ 0.07 AU. In the warm neutral medium, n ∼ 0.5 cm−3 , and
hence λ ∼ 2 × 1015 cm ∼ 100 AU ∼ 6 × 10−4 pc.
The situation is more complicated in a plasma. Consider a gas of fully
ionized hydrogen. The effective radius of interaction re for a free electron
can be found by setting the magnitude of its potential energy at a distance
re from an electron or proton equal to its thermal kinetic energy: is found
1
2 CHAPTER 1. FUNDAMENTALS OF GAS DYNAMICS
by setting
e2
∼ me ve2 . (1.3)
re
Since me ve2 ∼ kT , where T is the kinetic temperature of the free electrons,
we can write
e2 e2
re ∼ ∼ . (1.4)
me ve2 kT
The cross section is thus
πe4
σ ∼ πre2 ∼ , (1.5)
k2T 2
and the mean free path for an electron is
k2T 2
λ∼ . (1.6)
πe4 n
(Note: a more accurate calculation would contain the Coulomb logarithm
ln Λ, but this is good enough for an order-of-magnitude estimate.) In the hot
ionized interstellar medium, T ∼ 106 K and n ∼ 3 × 10−3 cm−3 . The mean
free path is thus λ ∼ 4 × 1019 cm ∼ 10 pc.
It’s useful at this point to provide a sketch of the different components of
the ISM. Molecular Clouds consist mainly of molecular gas (H2 , CO, etc.)
along with stubbornly solitary He atoms. Typical temperatures and number
densities are
T ∼ 15 K and n > ∼ 1000 cm .
−3
(1.7)
The Cold Neutral Medium consists mainly of neutral atomic gas (H I,
He I, etc.) Typical temperatures and number densities are
The Warm Neutral Medium consists mainly of neutral atomic gas, but
at a higher temperature and lower density than the Cold Neutral Medium.
Typical temperatures and number densities are
The Hot Ionized Medium consists of ionized gas at very high tempera-
tures. Typical temperatures and number densities are
T >
∼ 10 K and n <
6 −3
∼ 0.003 cm . (1.11)
Although the different components of the ISM are not in perfect pressure
equilibrium, it is notable that the pressures implied by the values of n and T
given above are all within an order of magnitude of each other, with a typical
value
P = nkT ∼ 3 × 10−13 dyne cm−2 ∼ 3 × 10−19 atm . (1.12)
The “atmosphere” (atm) is approximately equal to the pressure of the Earth’s
atmosphere at sea level. The ISM is obviously very low in pressure compared
to the air around us; thus, our intuitions about gas dynamics, which are
largely based on the behavior of air, may not always apply to the ISM.
The intergalactic medium (IGM) is typically at densities much lower than
even the Hot Ionized Medium of the ISM. The Warm/Hot Intergalactic
Medium (WHIM) has typical temperatures and number densities
1
The Boltzmann Equation also plays a key role in the dynamics of stellar systems.
5
The factor on the right hand side of the above equation is the rate at which
particles are bumped into a phase space volume by collisions with other par-
ticles. The word “collisions”, when used in this context, can be misleading,
since it implies that the individual particles behave like billiard balls, inter-
acting only when they touch each other. In an ionized gas, the individual
charged particles interact through long-range electromagnetic forces. A “col-
lision”, in that case, is a close encounter between charged particles during
which the electromagnetic force changes their direction of motion by ∼ 90◦
or more. Similarly, in stellar dynamics, a “collision” between stars is a close
encounter during which gravity changes their direction of motion by ∼ 90◦
or more.
We may rewrite equation (1.18) by making use of the fact that ⃗x˙ = ⃗v and
that ⃗v˙ = ⃗g , where ⃗g is the acceleration of a particle at position ⃗x with velocity
⃗v . The gravitational force acting on the particle results in an acceleration
⃗ ,
⃗g = −∇Φ (1.19)
where the potential Φ(⃗x) has contributions both from the self-gravity of the
system of particles and from any external gravitational fields that may be
present.2 The electromagnetic force acting on the particle results in an ac-
celeration # $
q ⃗ ⃗v ⃗ ,
⃗g = E+ ×B (1.20)
m c
⃗ is the electric field at
where q is the charge of the particle, m is its mass, E
⃗ is the magnetic field there.
the particle’s location, and B
The phase space continuity equation can now be written in the form
3
# $
∂f " ∂f ∂f df %%
+ vi + gi = % . (1.21)
∂t i=1 ∂xi ∂vi dt c
2
These notes are Newtonian. Unless I make an explicit statement to the contrary, all
motions are assumed to be non-relativistic, and gravity is assumed to be an inverse-square-
law force.
6 CHAPTER 1. FUNDAMENTALS OF GAS DYNAMICS
The partial differentials have been taken outside the integrals since ⃗x, ⃗v , and
t are all independent variables. The right hand side of the above equation
vanishes by the divergence theorem, as long as fm → 0 as v → ∞.
Then
∂ρ " ∂
+ (ρ⟨vi ⟩) = 0 . (1.25)
∂t i ∂xi
⃗u ≡ ⟨⃗v ⟩ . (1.26)
7
The velocity ⃗u is the bulk velocity at a given point in space. The velocity
⃗v of a particular particle may then be broken into two components:
⃗v = ⃗u + w
⃗ , (1.27)
It is useful to make this division because in some cases the viscous stress ten-
sor can be ignored. In other cases, the viscous stress tensor can be computed
in terms of the shear of the bulk velocity.
The equation for momentum conservation using the formalism that we
have adopted:
∂ " ∂
(ρuj ) + (ρui uj + P δij − πij ) = ρgj . (1.35)
∂t i ∂xi
The tensor form of the momentum conservation equation tells us that the
time derivative of a conserved quantity plus the divergence of a flux is equal
to a source term; this is the standard form of a conservation equation.
Let’s go one step further and look at the conservation of kinetic energy.
Multiply the collisionless Boltzmann equation by v 2 , and integrate over all
velocities.
∂ ! 2 3
" ∂ !
2 3
" ! ∂fm 3
v fm d v + vi v fm d v = − gi v 2 dv. (1.40)
∂t i ∂xi i ∂vi
9
Figure 1.1: Jim and Huckleberry Finn, drifting along, are Lagrangian ob-
servers; someone on the bank would be an Eulerian observer of the river.
By subtracting this equation from the equation for the conservation of the
total energy, we find the internal energy equation:
∂ ⃗ · (ρε⃗u) = −P ∇
⃗ · ⃗u − ∇
⃗ · F⃗ + Ψ ,
(ρε) + ∇ (1.48)
∂t
where
" ∂ui
Ψ≡ πij . (1.49)
i,j ∂xj
The function Ψ is the rate of viscous dissipation; viscosity converts bulk
kinetic energy into internal energy.
The partial time derivative of a function Q, which we have written as
∂Q/∂t, is the rate of change as viewed by an observer at a fixed coordinate
position, who is watching the gas flow by. Such an observer is called an
Eulerian observer. An alternative point of view is that of an observer who
is moving along with the bulk flow of the gas. Such an observer is called a
Lagrangian observer.
The Lagrangian time derivative,
DQ ∂Q ⃗ ,
= + ⃗u · ∇Q (1.50)
Dt ∂t
11
↔
The task at hand is to find expressions for the viscous stress tensor π and the
conductive heat flux F⃗ in terms of the density, pressure, and bulk velocity of
a gas. We will start with an ideal gas.
The distribution of random velocities for an ideal gas (otherwise known
as a “perfect gas”) is the Maxwellian distribution
mw2 3
13/2 # $
m
0
3
f (w)d
⃗ w= exp − dw, (2.1)
2πkT 2kT
where k is the Boltzmann constant, m is the particle mass, and T is the ki-
netic temperature. The mean square random velocity is then ⟨w2 ⟩ = 3kT /m.
Using equation (1.33), we find that the pressure is given by the familiar ideal
gas law
ρ
P = kT . (2.2)
m
For the Maxwellian distribution, ⟨wi wj ⟩ = 0 when i ̸= j; hence, all elements
↔
of π vanish. Furthermore, ⟨w2 wi ⟩ = 0, so the conductive heat flux also
vanishes. Ideal gases are delightfully simple: no viscosity, no heat conduction.
The conservation laws for an ideal gas take the simplified form
Dρ ⃗ · ⃗u
= −ρ∇ (2.3)
Dt
D⃗u 1⃗
= − ∇P + ⃗g (2.4)
Dt ρ
13
14 CHAPTER 2. VISCOSITY & HEAT CONDUCTION
Dε P⃗
= − ∇ · ⃗u . (2.5)
Dt ρ
These three equations are known collectively as the Euler equations. In
conjunction with the ideal gas law
3P 3kT
ε= = , (2.6)
2ρ 2m
they describe the motions of an inviscid, ideal gas of point masses.1 Com-
bining the continuity and energy conservation equations, we see that for an
ideal gas,
Dε P Dρ DV
= 2 = −P , (2.7)
Dt ρ Dt Dt
where V ≡ 1/ρ is the specific volume of the gas. The change in internal
energy of a ideal gas is equal to the P dV work that is done on the gas.
The first law of thermodynamics states that
T ds = dε + P dV , (2.8)
where T and P are the temperature and pressure of the gas, s is the specific
entropy, ε is the specific internal energy, and V is the specific volume. Since,
for an ideal gas,
Dε DV
= −P , (2.9)
Dt Dt
it is necessary that
Ds
T =0. (2.10)
Dt
An ideal gas, in the absence of heat sources and sinks, undergoes only adi-
abatic processes. (That is, its entropy remains constant.)
If a gas is compelled to have a constant volume (by being enclosed in a
rigid box, for instance), the first law of thermodynamics reduces to dq = dε,
where dq is the heat added. The specific heat at constant volume is then
# $ # $
∂q ∂ε
cV ≡ = . (2.11)
∂T V
∂T V
However, for an ideal gas, ε(V, T ) = ε(T ). Thus, we may write, quite gener-
ally,
dε = cV dT (2.12)
1
Inviscid = having no viscosity.
15
and
dq = cV dT + P dV . (2.13)
Now consider gas that is kept at a constant pressure. The ideal gas law tells
us that
k
P dV = dT (2.14)
m
when pressure is kept constant, and hence that
# $
∂q k
cP ≡ = cV + . (2.15)
∂T P
m
Note that cP > cV ; when pressure is held constant, some of the added heat
goes into P dV work instead of into internal energy.
The adiabatic index of a gas is defined as γ ≡ cP /cV . For an ideal gas
of point particles, ε = (3kT )/(2m), cV = (3k)/(2m), cP = (5k)/(2m) and
γ = 5/3. The adiabatic index for a gas of diatomic molecules is γ = 7/5.
The adiabatic index is a function of the number of degrees of freedom of
the particles; diatomic molecules have rotational degrees of freedom that
are not present for point masses. The internal energy ε, if the gas particles
are not spherical atoms, also includes rotational energy in addition to the
translational energy. In general, the internal energy is
1 kT 1 P
ε= = . (2.16)
γ−1 m γ−1 ρ
Now consider an adiabatic process, for which the first law of thermodynamics
may be written
dε + P dV = 0 (2.17)
P
cV dT − 2 dρ = 0 . (2.18)
ρ
For an ideal gas, # $
mP dP dρ
dT = − . (2.19)
kρ P ρ
Combining equations (2.18) and (2.19),
& # $ '
mP dP cV + k/m dρ
cV − =0. (2.20)
kρ P cV ρ
16 CHAPTER 2. VISCOSITY & HEAT CONDUCTION
where σ is the cross sectional area of the gas particles. To lowest order, the
coefficient of bulk viscosity, β, is equal to zero. Note that µ is independent
of the density of the gas. Note also that the viscosity of a gas increases
with temperature, in contrast to the behavior of liquids, in which viscosity
generally decreases with temperature.3 For neutral atomic hydrogen, the
coefficient of shear viscosity (including relevant factors of π) is
11/2
T
0
−3
µ = 6 × 10 poise . (2.28)
104 K
3
The common English simile, “as slow as molasses in January” makes reference to the
higher viscosity of molasses at lower temperatures.
18 CHAPTER 2. VISCOSITY & HEAT CONDUCTION
21
22 CHAPTER 3. SOUND & SHOCKS
ρ = ρ0 + ρ1 (x, t) (3.4)
u = u1 (x, t) (3.5)
P = P0 + P1 (x, t) (3.6)
g = g1 (x, t) . (3.7)
where |ρ1 |/ρ0 ≪ 1 and |P1 |/P0 ≪ 1. The linearized perturbation equations
are then
∂ρ1 ∂u1
+ ρ0 = 0 (3.8)
∂t ∂x
∂u1 dP ∂ρ1
ρ0 + | = g1 ρ 0 (3.9)
∂t dρ 0 ∂x
∂g1
= −4πGρ1 . (3.10)
∂x
In writing the equation of motion, I have made the implicit assumption that
P = P (ρ). Taking the time derivative of equation (3.8), and subtracting the
spatial derivative of equation (3.9), we find
∂ 2 ρ1 dP ∂ 2 ρ1
− | = 4πGρ0 ρ1 . (3.11)
∂t2 dρ 0 ∂x2
The density and pressure perturbations that propagate through the medium
with velocity ±a are sound waves. For a polytrope with adiabatic index γ,
the sound speed at density ρ0 and pressure P0 is
# $1/2 # $1/2
γP0 γk
a0 = = T . (3.13)
ρ0 m
For a neutral atomic gas, the sound speed is a = 0.12 km s−1 µ−1/2
a (T /1 K)1/2 ,
where µa is the mean atomic mass in units of the proton mass.
23
Using these relations in conjunction with the continuity equation, the energy
equation takes the form
∂ ∂2
(P1 − a20 ρ1 ) = γχ 2 (P1 − a20 ρ1 /γ) , (3.23)
∂t ∂x
where χ = K/(ρ0 cP ).
If the perturbations are sine waves, of the form
1 − ik 2 χ/ω 2 2
ω2 = a k + iµ′ k 2 ω/ρ0 . (3.30)
1 − iγk 2 χ/ω 0
ρ 1 u 1 = ρ 2 u2 (3.38)
ρ1 u21+ P1 = ρ2 u22 + P2 (3.39)
1 2 1 2
u1 + ε1 + P1 /ρ1 = u + ε2 + P2 /ρ2 . (3.40)
2 2 2
28 CHAPTER 3. SOUND & SHOCKS
Figure 3.2: The geometry of a plane parallel shock; we are in the shock’s
frame of reference, with pre-shocked gas flowing in from the left.
29
I have made the implicit assumption that γ is the same for the preshock and
postshock gas.
The Rankine-Hugoniot conditions are just the conservation equations in
a new guise. A dimensionless number that is often cited in the context of
shocks is the Mach number,
$1/2
ρ1 u21
#
M1 ≡ u1 /a1 = . (3.42)
γP1
The Mach number is the ratio of the velocity of the shock (relative to the
unshocked medium) to the sound speed in the unshocked medium. Using
the Rankine-Hugoniot conditions, we may solve for the density, pressure,
and temperature jumps in terms of M1 .
ρ2 (γ + 1)M12 u1
= = (3.43)
ρ1 (γ − 1)M12 + 2 u2
2
P2 2γM1 − (γ − 1)
= (3.44)
P1 γ+1
T2 [(γ − 1)M12 + 2][2γM12 − (γ − 1)]
= . (3.45)
T1 (γ + 1)2 M12
Thus, no matter how strong the shock is, the ratio ρ2 /ρ1 has a finite value;
for a monatomic gas, with γ = 5/3, the ratio is ρ2 /ρ1 = 4. However, a strong
30 CHAPTER 3. SOUND & SHOCKS
shock is very efficient at converting the bulk kinetic energy of the upstream
gas (∼ ρ1 u21 ) into thermal energy.
A weak shock is defined as one that is barely supersonic, with M1 − 1 =
ϵ ≪ 1. For a weak shock,
ρ2 4
≈ 1+ ϵ (3.49)
ρ1 γ+1
P2 4γ
≈ 1+ ϵ (3.50)
P1 γ+1
T2 4(γ − 1)
≈ 1+ ϵ. (3.51)
T1 γ+1
Generally speaking, a shock converts supersonic gas (M1 > 1) into subsonic
gas (in the shock’s frame of reference). It increases density, decreases bulk
velocity (relative to the shock front), increases pressure, and increases tem-
perature.
The conversion of bulk kinetic energy to random thermal energy occurs
by dissipation within the shock layer itself. Within the shock layer, a jump
in velocity of ∆u ∼ a1 ∼ (kT /m)1/2 occurs over a length ∆x ∼ λ ∼ 1/(nσ).
The xx component of the viscous stress tensor within the shock is
√ √
4 ∆u mkT kT
πxx = ( π + β) ∼ √ (nσ) ∼ nkT . (3.52)
3 ∆x σ m
The increase in the specific entropy as the gas crosses the shock front is
s2 − s1 = cP ln(T2 /T1 ) − (k/m) ln(P2 /P1 ). For an extremely strong shock
(M1 → ∞), the entropy increase is s2 − s1 ∝ ln M1 .
So far, we’ve been dealing with normal shocks; that is, shocks in which
the the velocity vector ⃗u1 of the upstream, unshocked gas is perpendicular
to the shock front. However, some shocks are better described as oblique
shocks, in which ⃗u1 approaches the shock front at an angle other than 90
degrees. For instance, the shock waves associated with spiral arms of galaxies
can be approximated as oblique shocks. An oblique shock is illustrated in
Figure 3.3. Again, let us put ourselves into a frame of reference that is
moving along with a plane parallel shock front. This time, however, the
unshocked gas is flowing into the shock at an angle φ with respect to the
plane of the shock front. Thus, the bulk velocity u1 can be broken down into
a perpendicular component
u⊥1 = u1 sin φ (3.53)
31
Figure 3.3: The geometry of an oblique shock; we are in the shock’s frame
of reference, with pre-shocked gas flowing in from the left.
32 CHAPTER 3. SOUND & SHOCKS
which occurs when tan2 φ = (γ + 1)/(γ − 1). For a monatomic gas (γ = 5/3)
the maximum angle of refraction is ψm = 36.9◦ , which occurs when φ = 63.4◦ .
Even an arbitrarily strong shock can’t divert the flow through an angle of 90◦ .
If a blunt object, therefore, moves through a gas at supersonic speeds, then
a detached bow shock must form ahead of the object, as shown in Figure 3.4.
34 CHAPTER 3. SOUND & SHOCKS
Chapter 4
Radiative &
Magnetohydrodynamic Shocks
I have been dealing, so far, with non-radiative shocks. Since, as we have seen,
a shock raises the density and temperature of the gas, it is quite likely, under
astrophysical conditions, that the post-shock gas will be highly radiative.
Consider, again, plane parallel shocks in which the gas flow is perpen-
dicular to the shock front, as shown in Figure 4.1. A gas with density ρ1
and pressure P1 flows into the shock front with a velocity u1 relative to the
front. The pre-shock gas is assumed to be in thermal equilibrium, and is not
radiating energy. The gas passes through the thin shock layer; immediately
downstream, it has density ρ2 , pressure P2 , and bulk velocity u2 given by
the Rankine-Hugoniot jump conditions. So far, everything is the same as in
the nonradiative shock. Now, however, we assume that there is an optically
thin radiative relaxation layer downstream of the shock front, in which
the cooling function L(ρ, T ) is greater than zero.1
Within the radiative relaxation layer, the conservation equations are
ρu = constant (4.1)
2
ρu + P = constant (4.2)
dε du
ρu = −P − ρL . (4.3)
dx dx
For an ideal gas, the energy equation reduces to
u dP γ du
+ P = −ρL (4.4)
γ − 1 dx γ − 1 dx
1
The cooling function L has units of energy per unit mass per unit time.
35
36 CHAPTER 4. RADIATIVE & MHD SHOCKS
Figure 4.1: The geometry of a radiative plane parallel shock; we are in the
shock’s frame of reference, with pre-shocked gas flowing in from the left.
37
ρu = ρ2 u2 (4.6)
ρu + P (x) = ρ2 u22 + P2
2
(4.7)
1 du
(γP/ρ − u2 ) = −L(ρ, T ) (4.8)
γ−1 dx
mP
T = (4.9)
k ρ
tell us the physical conditions within the radiative relaxation layer. Since
the post-shock gas is subsonic (u2 < γP/ρ) and since L > 0, equation (4.8)
indicates that the flow velocity u decreases as you move through the radiative
relaxation layer away from the shock front. Since u decreases, equation (4.6)
tells us that ρ increases. From equations (4.6) and (4.7), the pressure is
For a strong shock P2 > ρ2 u22 , and the pressure increases only slightly as the
gas flows through the radiative relaxation layer. The temperature, in terms
of u, is
ρ2 u22
! "# $
u u
! "
T = T2 1+ (1 − u/u2 ) ≈ T2 . (4.11)
u2 P2 u2
For a strong shock, T ∝ u and the temperature decreases in the radiative
relaxation layer. The gas will approach a steady state equilibrium in which
L = 0, and the density, pressure, temperature, and bulk velocity have the
constant values ρ3 , P3 , T3 , and u3 .
38 CHAPTER 4. RADIATIVE & MHD SHOCKS
ρ 3 u 3 = ρ 1 u1 (4.12)
ρ3 u23 + P3 = ρ1 u21 + P1 (4.13)
T3 = T1 . (4.14)
Making use of the equation of state P = ρkT /m, and using the “isothermal
sound speed” aT ≡ (kT1 /m)1/2 , we find that the increase in density is
"2
ρ3 u1
!
= = MT2 , (4.15)
ρ1 aT
⃗ ·E
∇ ⃗ = 4πρe (4.16)
39
Figure 4.2: The density, bulk velocity, temperature, and pressure of material
passing through a Mach 2 isothermal shock.
40 CHAPTER 4. RADIATIVE & MHD SHOCKS
⃗
⃗ ×E
∇ ⃗ = − 1 ∂B (4.17)
c ∂t
⃗ ·B
∇ ⃗ = 0 (4.18)
⃗
∇ ⃗ = 4π ⃗je + 1 ∂ E .
⃗ ×B (4.19)
c c ∂t
In the above equations, ρe is the charge density and ⃗je is the current density.
In nonrelativistic systems, the term c−1 ∂ E/∂t⃗ in equation (4.19) may be
ignored.
In an ionized gas, the current, electric field, and magnetic flux are related
through the expression
E⃗ = 1 ⃗je − 1 ⃗u × B⃗ , (4.20)
σ c
where σ is the electrical conductivity. Using equation (4.19) to eliminate je ,
⃗ = c ⃗ ⃗ − 1 ⃗u × B
⃗ .
E ∇×B (4.21)
4πσ c
Substituting this result into equation (4.17), we find the basic equation for
⃗
the evolution of B:
⃗
∂B ⃗ × (B
+∇ ⃗ × ⃗u) = −∇
⃗ × (η ∇
⃗ × B)
⃗ . (4.22)
∂t
The quantity η is the electrical resistivity,
c2
η≡ . (4.23)
4πσ
If η is constant throughout the gas, equation (4.23) may be rewritten in the
form
⃗
∂B
+∇⃗ × (B ⃗ × ⃗u) = η∇2 B
⃗ , (4.24)
∂t
where I have made use of the fact that ∇ ⃗ ·B ⃗ = 0. If ⃗u = 0, then the
⃗ with a diffusion
above equation represents the diffusion of the vector field B,
constant η.
The electric and magnetic fields exert force on the charged particles that
make up an ionized gas. For a completely ionized gas, with ions of charge
Ze, the mean force per unit volume is
⃗ + 1 ⃗ui × B)
f⃗L = Zeni (E ⃗ + 1 ⃗ue × B)
⃗ − ene (E ⃗ . (4.25)
c c
41
The ions have a number density ni and a bulk velocity ⃗ui ; the electrons
have a number density ne and a bulk velocity ⃗ue . Since maintaining charge
neutrality requires that Zeni − ene = 0, and the current density is given by
the relation ⃗je = Zeni⃗ui − ene⃗ue , the Lorenz force per unit volume is
1
f⃗L = ⃗je × B
⃗ (4.26)
c
1 ⃗ ⃗ ×B
⃗ .
= (∇ × B) (4.27)
4π
Using the appropriate identity from vector algebra,
1 ⃗ ⃗ ⃗ ⃗ B2
% &
f⃗L = (B · ∇)B − ∇ . (4.28)
4π 8π
The first term on the right hand side represents a ‘magnetic tension’; it
exists only if the magnetic field lines are curved, and acts in such a way as
to straighten them out. The second term on the right hand side represents
the gradient of the ‘magnetic pressure’, Pm = B 2 /(8π).
The magnetic field also adds a heating term to the internal energy con-
servation equation. An electric current je passing through a medium of re-
sistivity η converts electromagnetic energy to heat energy at the rate
4πη 2 η ⃗ ⃗ 2.
P = 2
je = |∇ × B| (4.29)
c 4π
For the sake of monumental completeness, let’s write down the basic equa-
tions of MHD in all their glory.
Dρ ⃗ · ⃗u
= −ρ∇ (4.30)
Dt
D⃗u ⃗ + ∇· ⃗ ↔ ⃗ + f⃗rad + 1 (∇⃗ × B)
⃗ ×B ⃗
ρ = −∇P π −ρ∇Φ (4.31)
Dt 4π
Dε ⃗ · F⃗ + ψ + Γ − Λ + η |∇
⃗ · ⃗u − ∇ ⃗ × B|
⃗ 2
ρ = −P ∇ (4.32)
Dt 4π
∇2 Φ = 4πGρ (4.33)
⃗
∂B ⃗ × (B
+ ∇ ⃗ × ⃗u) = −∇
⃗ × (η ∇
⃗ × B)
⃗ (4.34)
∂t
⃗ ·B
∇ ⃗ = 0 (4.35)
1 P
ε = . (4.36)
γ−1 ρ
42 CHAPTER 4. RADIATIVE & MHD SHOCKS
Now that we’ve taken the trouble to write down these complicated equations,
we will ruthlessly simplify them. First we will ignore viscosity and heat
↔
conduction, setting π= 0, ψ = 0, and F⃗ = 0. We will ignore gravity, and set
⃗ = 0. We will ignore radiative effects, and set f⃗rad = 0 and Γ − Λ = 0.
∇Φ
We will ignore diffusion of the magnetic field, and set η = 0.
All of these simplifications will make it possible for us to consider a steady-
state planar shock in the presence of a magnetic field. If a planar shock front
exists, then the velocity vector ⃗u and the magnetic field vector B ⃗ can both be
broken down into components perpendicular to the shock front (designated
by the subscript ⊥) and parallel to the shock front (designated by the sub-
script ∥). Quantities measured just upstream of the shock will be designated
by the subscript ‘1’, as usual, and quantities measured just downstream of
the shock will be designated by the subscript ‘2’. The jump conditions across
the shock are:
ρ1 u⊥1 = ρ2 u⊥2 (4.37)
2 2
2
B∥1 B∥2
ρ1 u⊥1 + P1 + = ρ2 u2⊥2 + P2 + (4.38)
8π 8π
B⊥1 B∥1 B⊥2 B∥2
ρ1 u⊥1 u∥1 − = ρ2 u⊥2 u∥2 − (4.39)
4π 4π
γ P1 u21 B∥1
ρ1 u⊥1 ( + ) − (B⊥1 u∥1 − B∥1 u⊥1 ) = (4.40)
γ − 1 ρ1 2 4π
γ P2 u22 B∥2
ρ2 u⊥2 ( + )− (B⊥2 u∥2 − B∥2 u⊥2 )
γ − 1 ρ2 2 4π
B⊥1 u∥1 − B∥1 u⊥1 = B⊥2 u∥2 − B∥2 u⊥2 (4.41)
B⊥1 = B⊥2 (4.42)
Note that the component of the velocity parallel to the shock is no longer
conserved. The discontinuity in u∥ occurs because there is a current within
the shock front of strength (c/4π)(B∥2 − B∥1 ), which increases u∥ by a factor
B⊥
u∥2 − u∥1 = (B∥2 − B∥1 ) . (4.43)
4πρu⊥
The most mathematically tractable MHD shock is a normal shock (u∥ = 0,
u⊥ = u) in which the magnetic field is parallel to the shock front (B⊥ = 0,
B∥ = B). In this case, the jump relations simplify to the form
ρ 1 u 1 = ρ 2 u2 (4.44)
43
B12 B2
= ρ2 u22 + P2 + 2
ρ1 u21 + P1 + (4.45)
8π 8π
γ P1 1 2 B1 γ P2 1 2 B2
+ u1 + = + u2 + 2 (4.46)
γ − 1 ρ1 2 4πρ1 γ − 1 ρ2 2 4πρ2
B1 u1 = B2 u2 (4.47)
Using these four equations, we can solve for the density jump ρ2 /ρ1 in terms
of the upstream Mach number M1 and the ratio of the magnetic pressure to
the thermal pressure,
B12
α1 ≡ . (4.48)
8πP1
After discarding the trivial solution ρ2 = ρ1 , the four jump relations simplify
to the quadratic relation
% &2 % &
ρ2 ρ2
2(2 − γ)α1 + γ[(γ − 1)M12 + 2(α1 + 1)] − γ(γ + 1)M12 = 0 . (4.49)
ρ1 ρ1
When the magnetic pressure is relatively insignificant (α1 ≪ M12 ) the change
in density is approximately
γ + M12
# $
ρ2 4
= y0 1 − α1 , (4.50)
ρ1 γ [2 + (γ − 1)M12 ]2
where y0 is the density ratio in the absence of a magnetic field. The presence
of a magnetic field thus tends to decrease the density jump from what it
would be in the absence of magnetism. In fact, if we again examine the
quadratic equation for ρ2 /ρ1 , we see that ρ2 > ρ1 when
2
M12 > Mcr2 ≡ 1 + α1 . (4.51)
γ
The existence of a magnetic field permits you to have supersonic motions
(M > 1) without the formation of shocks. Magnetosonic waves (fluctua-
tions in the combined magnetic and gas pressure) can travel ahead of the
shock front. Fluctuations in the gas pressure have a characteristic speed
a ∝ (P/ρ)1/2 . Similarly, fluctuations in the magnetic pressure have a char-
acteristic speed, called the Alfven velocity, uA ≡ B/(4πρ)1/2 ∝ (Pm /ρ)1/2 .
Magnetosonic waves that travel in a direction perpendicular to B ⃗ have a
2 2 1/2 1/2
phase velocity u = (a + uA ) = a[1 + α(2/γ)] .
44 CHAPTER 4. RADIATIVE & MHD SHOCKS
Chapter 5
One way to create a shock is to inject a very large amount of energy into
a very small volume. This happens in the interstellar medium whenever a
supernova goes off. Suppose an explosion instantaneously injects an amount
of energy E into an ambient medium of uniform density ρ1 . The initial energy
release is considered to take place within an infinitesimally small volume.
Afterward, however, a spherical shock front will expand into the ambient
medium. Early in the course of expansion, the pressure within the shock,
P2 ∼ ρ1 u2sh , is much larger than the ambient pressure P1 and any radiated
energy is much smaller than the explosion energy E. This regime, during
which the energy E remains constant, is known as the blastwave regime. In
a blastwave, the expansion velocity ush (r, t), density ρ(r, t), pressure P (r, t),
and other properties, are determined solely by the two initial parameters of
the system, E and ρ1 .
The energy E has the dimensionality M L2 T −2 ; the density ρ1 has the
dimensionality M L−3 . These two parameters cannot be combined to form
a characteristic length scale or time scale for the problem. The solution for
the expanding shock front must then be a scale-free or self-similar solution.
The self-similar solution is a function of the dimensionless variable ξ, where
ξ ≡ rtl ρm n
1 E ; (5.1)
45
46 CHAPTER 5. SPHERICAL BLASTWAVES & SNRS
that the required solution has the exponents l = −2/5, m = 1/5, and n =
−1/5, or
ρ1 1/5
! "
ξ=r . (5.2)
Et2
When expressed in the dimensionless units, the properties of the expanding
shock front will depend only on ξ. For instance, the radius of the spherical
shock is $1/5
Et2
#
rsh = ξ0 , (5.3)
ρ1
where ξ0 is a factor of order unity (for γ = 5/3, it turns out that ξ0 = 1.17).
The rate of expansion of the shock is
# $1/5
2 E
ush = ξ0 . (5.4)
5 ρ1 t3
Thus,
# $1/2
2 5/2 E −3/2
ush = ξ0 rsh . (5.5)
5 ρ1
The expanding shock wave slows as it expands. Using typical values for
supernova explosions,
# $1/5 # $−1/5 # $2/5
E ρ1 t
rsh = 2.3 pc 51
(5.6)
10 erg 10 g cm−3
−24 100 yr
# $1/5 # $−1/5 # $−3/5
E ρ1 t
ush = 9000 km s−1 51
(5.7)
.
10 erg 10 g cm−3
−24 100 yr
For instance, SN1054 (which gave birth to the Crab Nebula) should have a
radius rsh ∼ 5 pc and expansion velocity ush ∼ 2000 km s−1 , if it is still in the
blastwave phase of expansion. In fact, the Crab Nebula is observed to have
a radius of 3 pc and an expansion velocity of 900 km s−1 . In fact, the Crab is
not exactly spherical and its expansion is not exactly energy-conserving, so
the discrepancy is not surprising.
Obviously, the self-similar blastwave solution is not good for all times.
As t → 0, for instance, the solution predicts infinitely rapid expansion. The
self-similar blastwave solution breaks down when the predicted expansion
velocity is larger than the velocity with which the supernova initially ejects
matter into the ambient medium. For a typical supernova, the initial ejection
47
velocity is uej ∼ 104 km s−1 ; thus, the blastwave solution is invalid when the
supernova remnant is younger than ∼ 70 yr, for our ‘standard’ supernova
parameters.
The self-similar blastwave solution also breaks down when the expanding
shock front is very old. The pressure immediately inside the shock is, for an
extremely strong shock,
2
P2 = ρ1 u2sh (5.8)
γ+1
8 1 5 −3
= ξ Er . (5.9)
25 γ + 1 0 sh
Thus, the pressure within the blastwave falls off rapidly with radius. If we
insert numerical values, we find that
# $1/3 # $−1/3
E P2
rsh ≈ 300 pc 51
. (5.10)
10 erg 4 × 10 dyne cm−2
−13
Thus, when a typical supernova remnant reaches a radius of 300 parsecs, the
internal pressure has dropped until it is comparable to the ambient pressure of
the interstellar medium. This violates our initial assumption that P2 ≫ P1 .
For our typical supernova remnant, the assumption of negligible pressure
breaks down badly at radii of rsh ∼ 300 pc, corresponding to an age of t ∼
30 Myr.
In finding the blastwave solution, however, we made another assumption;
namely, that the energy lost in radiation was small compared to the explosion
energy E. At what point does this assumption break down? The temperature
within the shock will be, for ionized matter of cosmic abundances,
2(γ − 1) m 2
T2 = u (5.11)
(γ + 1)2 k sh
# $# $2/5 # $−2/5 # $−6/5
5 m E ρ1 t
≈ 2 × 10 K (5.12).
10−24 g 51
10 erg 10 g cm−3
−24 105 yr
At an age of 105 yr, the radius of the shock will be rsh ∼ 30 pc. The tem-
perature immediately within the shock will be T2 ∼ 2 × 105 K; the cooling
function at this temperature has the value ρL ∼ 10−21 erg cm−3 s−1 . The
total energy radiated during the course of 105 years will be, very roughly,
4π 3
Erad ∼ ρL( rsh )t ∼ 1051 erg . (5.13)
3
48 CHAPTER 5. SPHERICAL BLASTWAVES & SNRS
Figure 5.1: An expanding supernova remnant makes the transition from the
blastwave phase (upper image) to the snowplow phase (lower image). [Shu,
The Physics of Astrophysics, Volume II, Figure 17.4]
50 CHAPTER 5. SPHERICAL BLASTWAVES & SNRS
The radius of the shock, for the adiabatic blastwave, is rsh = ξo (Et2 /ρ1 )1/5 ,
and the velocity of expansion is ush = (2/5)rsh /t. If the shock is very strong,
the density, bulk velocity and pressure immediately inside the shock front
are found from the Rankine-Hugoniot conditions:
γ+1
ρ2 = ρ1 (5.17)
γ−1
2
u2 = ush (5.18)
γ+1
2
P2 = ρ1 u2sh . (5.19)
γ+1
But now, we ask, what are ρ, u, and P for 0 < r < rsh ?
We can find these quantities by solving the equation of mass conservation,
∂ρ 1 ∂
+ 2 (r2 ρu) = 0, (5.20)
∂t r ∂r
the equation of momentum conservation,
∂u ∂u ∂P
+u =− , (5.21)
∂t ∂r ∂r
and the equation of energy conservation,
∂ 1 1 ∂ 1
(ρε + ρu2 ) + 2 [r2 ρu(ε + P/ρ + u2 )] = 0 , (5.22)
∂t 2 r ∂r 2
where ε = P/[(γ − 1)ρ].
Since the solutions for the blastwave is self-similar, the solutions must
take the form
% &
γ+1
ρ(r, t) = ρ1 α(ξ) (5.23)
γ−1
% &
4 r
u(r, t) = v(ξ) (5.24)
5(γ + 1) t
r2
% &
8
P (r, t) = ρ1 2 p(ξ) . (5.25)
25(γ + 1) t
The factors of ρ1 , r, and t provide the proper dimensionality for the solutions.
The numerical factors in square brackets are inserted so that the dimension-
less density, velocity, and pressure have the normalization α(ξ0 ) = v(ξ0 ) =
51
Figure 5.2: The Sedov solution for a spherical blastwave with γ = 5/3, in
units of the immediate post-shock values. [Shu, The Physics of Astrophysics,
Volume II, Figure 17.3]
The time has come to compare our theoretical results with observations
of actual supernova remnants (SNRs). Supernova remnants can be observed
that correspond to the first three phases of theoretical expansion. The first
phase is the free expansion of the matter of the exploded star. During this
phase, matter moves outward with a constant velocity veject . The remnant
associated with SN1987A is in the free expansion phase. The remnant Cas
A is at the end of its free expansion phase; it is ∼ 300 years old. At visible
wavelengths, many ‘knots’ of emission, moving radially outward with a ve-
locity of ∼ 6000 km s−1 , can be seen. The knots are rich in oxygen, and are
interpreted as being clumps of matter that have been ejected from the center
of the star, where nucleosynthesis took place.
When the amount of interstellar gas swept up becomes comparable to
the initial mass of ejected matter, the SNR enters the blastwave phase. The
remnant of Tycho’s supernova is in the blastwave phase. The Crab nebula is
also in the blastwave phase; its age is ∼ 940 yr, and its expansion velocity is
∼ 900 km s−1 .
When the energy radiated just behind the shock front is comparable to the
initial energy of the explosion, the SNR enters the snowplow phase. During
this phase, a dense cool shell forms directly behind the shock front. In theory,
the momentum of this shell is conserved. Observationally, it is found that
the internal pressure of SNRs in the snowplow phase is large enough to give
a significant push to the shell. The Cygnus Loop (alias the Veil Nebula) is
in the snowplow phase; its age is ∼ 4 × 104 yr, and its expansion velocity
is ∼ 120 km s−1 . The dense shell is thermally unstable to the formation of
filaments; such filaments can be seen in optical photographs of the Cygnus
Loop.
Although the stages of SNR evolution can be fairly well described by our
simple theory, there are many deviations that should be mentioned. First of
all, the progenitors of SNRs are massive hot stars that have copious stellar
winds. Thus, at the time that the supernova goes off, it will be surrounded
by a low-density bubble that has been excavated by the stellar wind. As
long as the supernova ejecta are expanding into this low-density region, they
will sweep up very little mass, and the free expansion phase will be pro-
longed. Another complication is that the ambient interstellar medium is not
uniform. This inhomogeneity will blur the distinction between the blastwave
and snowplow phases; a portion of the shock front that is passing through a
dense cloud may be in the snowplow phase, while a neighboring portion that
is passing through the rarefied intercloud medium is still in the blastwave
53
phase.
It should also be noted that massive stars tend to be born in stellar
associations. They enter their supernova phase before they have a chance
to drift apart. Thus, the remnants of the neighboring supernovas will merge
to form a single superbubble, which may be hundreds of parsecs across.
Such superbubbles are a primary source of the hot ionized component of the
interstellar medium.
54 CHAPTER 5. SPHERICAL BLASTWAVES & SNRS
Chapter 6
A shock front is the abrupt transition between two regions with different
densities, bulk velocities, and pressures. An ionization front is the abrupt
transition between a region of ionized gas and a region of neutral gas. The
boundary of an HII region is an approximation to a spherical ionization
front.
An HII region is a volume of photoionized gas surrounding a hot young
star (of spectral type O or B). A planetary nebula also consists of a volume
of ionized gas surrounding a central photoionizing star; the central star in a
planetary nebula, however, is an evolved star approaching its death throes.
Planetary nebulae tend to be denser and smaller than HII regions. The
physics of planetary nebulae is complicated by the fact that the photoioniza-
tion phase is preceded by the ejection of a large quantity of neutral gas with
a velocity of ∼ 20 km s−1 . Because the structure of an HII region is simpler,
I will use it as my basic example of a photoionization region.
The gas in an HII region is ionized by photons emitted by the central star.
The ionization energy of hydrogen is 13.6 eV, corresponding to a wavelength
of 912 angstroms. To photoionize significant amounts of hydrogen, the stellar
temperature must be T > ∼ 25,000 K. As the star continues to emit photons,
the HII region becomes larger. Eventually, enough energy is pumped into
the ionized material to raise its temperature and cause it to expand. Thus, a
generic HII region consists of a low-density, high-temperature, ionized region
expanding outward into the ambient high-density, low-temperature, neutral
medium.
55
56 CHAPTER 6. IONIZATION FRONTS
The original classic study of HII regions was performed by Bengt Strömgren
in 1939. He considered a simple model, which still managed to convey the
basic physics of HII regions. We will examine the ‘Strömgren sphere’, as his
model is called, and later examine real HII regions to see how they differ
from the ideal case. Strömgren assumed the presence of a uniform medium,
consisting of neutral atomic hydrogen, with number density n0 . Suddenly,
at time t = 0, a hot star turns on at the origin. The star’s total output
of ionizing photons (in photons per second) is Nu . The initial effect of the
star’s radiation will be solely to ionize the hydrogen. Thus, the star will
be surrounded by a sphere of electrons and protons, with number density
ne = np = n0 . The spherical volume of ionized gas will be separated from
the neutral medium by a thin transition layer, whose thickness is comparable
to the mean free path of an ionizing photon in the neutral medium. The mean
free path λi is given by the relation λi = 1/(n0 σi ), where the cross-section
for photoionization is σi ≈ 6 × 10−18 cm2 for hydrogen atoms.
Initially, the HII region expands outward at a rate given by the relation
dR
Nu = n0 4πR2 , (6.1)
dt
which integrates to
3Nu 1/3
! "
R(t) = t . (6.2)
4πn0
However, an ionized hydrogen atom does not remain ionized forever. The
electrons and protons collide and recombine to form neutral hydrogen at a
rate Nrec = αne ni . The recombination rate α(T ) has a value of α = 2.6 ×
10−13 cm3 s−1 at a temperature T = 104 K, which is a typical temperature
for an HII region. Once recombinations become important, the radius of the
HII region is given by the relation
dR 4π
Nu = 4πn0 R2 + αn20 R3 . (6.3)
dt 3
This has the solution
' (1/3
R(t) = Rs 1 − e−t/ts , (6.4)
where the characteristic time scale is ts = 1/(αn0 ) and the characteristic
length scale is the Strömgren radius,
# $1/3
3Nu
Rs ≡ . (6.5)
4παn20
57
where
)
uR ≡ a 2 + a22 − a21 (6.12)
)
uD ≡ a 2 − a22 − a21 . (6.13)
Physical reality demands that the ratio ρ2 /ρ1 be a real number. Thus, the
velocity of the ionization front (relative to the neutral gas) must have either
u1 ≥ uR or u1 ≤ uD . The rapidly propagating ionization fronts (u1 ≥ uR ) are
called R-type fronts; the slowly propagating ionization fronts (u1 ≤ uD )
are called D-type fronts. In this naming scheme, ‘R’ stands for ‘rarefied’
and ‘D’ stands for ‘dense’.1
In HII regions, it’s a fair approximation that a2 ≫ a1 , so uR ≈ 2a2 and
uD ≈ (a1 /a2 )2 a2 /2. An R-type front always has u1 > uR > a2 > a1 , and is
supersonic with respect to the neutral medium. A D-type front always has
u1 < uD < a1 < a2 , and is subsonic with respect to the neutral medium.
For a given propagation velocity u1 , there are two possible values of ρ2 /ρ1 ,
corresponding to taking the plus or the minus sign in the solution of the
quadratic equation. This is shown graphically in Figure 6.2, which is a plot
of the ratio ρ2 /ρ1 as a function of u1 . A front that has the larger density
contrast is called a strong front; a front that has the smaller density contrast
is a weak front. Thus, there are four types of ionization front possible; a
weak R-type front, a strong R-type front, a weak D-type front, and a strong
D-type front. It turns out that strong R-type fronts are unstable, and are not
seen in nature. Weak R-type fronts and both strong and weak D-type fronts
are known to exist, however, around HII regions and planetary nebulae.
In a weak R-type front, the incoming neutral gas is supersonic with re-
spect to the front (u1 > a1 ), and the outflowing ionized gas is also supersonic
with respect to the front (u2 > a2 ). In a strong R-type front, the incoming
neutral gas is supersonic, and the outflowing ionized gas is subsonic. In a
weak D-type front, the incoming neutral gas is subsonic, and the outflowing
ionized gas is also subsonic. In a strong D-type front, the incoming gas is
subsonic, and the outflowing ionized gas is supersonic.
Consider, for example, an expanding HII region. When the ionizing star
is first ‘turned on’, the spherical ionization front is very small; hence the flux
Φi is very large, resulting in a high velocity u1 . Thus, the front is initially
1
For a given flux Φi , as the gas becomes more rarefied (ρ1 → 0), the velocity u1 → ∞,
and an R-type front results. Similarly, as ρ1 → ∞, u1 → 0, and a D-type front results.
60 CHAPTER 6. IONIZATION FRONTS
Figure 6.2: Physically permissible values of the density contrast ρ2 /ρ1 across
an ionization front with a2 = 10a1 .
61
ρ2 a2
≈ 1 + 22 weak R . (6.14)
ρ1 u1
A weak R-type front compresses the gas only slightly. This period, when
the interior and exterior densities are nearly the same, is when Strömgren’s
approximation holds good.
As the ionization front surrounding the HII region expands, the flux of
ionizing photons Φi steadily decreases. Eventually, the velocity u1 drops to
a value u1 = uR . At this point, the front is an ‘R-critical’ front; the density
ratio is ρ2 /ρ1 ≈ 2 and the velocity of the ionized gas is u2 ≈ a2 . Once
the ionization front slows still farther, the R-type front can no longer exist.
What happens next is illustrated in Figure 6.3. When the velocity u1 drops
below uR , the R-critical front splits into a shock front followed by a D-critical
front. The shock front increases the density of the gas (by a factor of 4 if
M1 ≫ 1 and γ = 5/3), decreases the velocity of the gas, and increases the
sound speed. Because of the decrease in the bulk velocity and the increase in
the sound speed, the flow of neutral gas is now subsonic, and is ready to pass
through a D-type front that ionizes it. As the HII region expands farther,
the leading shock front gradually weakens, and the trailing D-critical front
develops into a weak D-type front. Thus, a bit of neutral gas will be first
compressed by the shock and then blasted with UV photons and ionized.
The shock front expands outward and decreases in amplitude until it fi-
nally becomes a sound wave of infinitesimal amplitude expanding outward
with velocity u1 = a1 . If the photoionizing star is immortal, then the ion-
ization front will expand more and more slowly until HII region attains an
equilibrium state in which the pressure of the expanded ionized gas is equal
to the pressure of the neutral surrounding medium. The final equilibrium
radius, Rfinal , is given by the relation
# $1/3
3Nu
Rfinal = , (6.15)
4παn2final
where the density nfinal within the HII region is given by the requirement of
pressure equilibrium:
2nfinal T2 = n0 T1 . (6.16)
62 CHAPTER 6. IONIZATION FRONTS
Figure 6.3: The transition from an R-critical front to a shock front followed
by a D-critical front.
63
Thus,
"2/3
2T2
!
Rfinal = Rs ∼ 27Rs . (6.17)
T1
In practice, however, HII regions never reach this equilibrium state. The
massive O and B stars that ionize them become supernovae before then.
Observers classify real HII regions into five categories that correlate well
with the age of the region. An ultracompact HII region contains a single,
very young, hot star. Ultracompact HII regions are embedded deep within
molecular clouds, and can be observed only in the radio and the infrared.
They have radii R < ∼ 0.1 pc and densities n >
4 −3
∼ 10 cm . An example of
an ultracompact HII region is the Becklin-Neugebauer object, located in a
molecular cloud near the Orion nebula.
As the HII region expands, it becomes a compact HII region. A compact
HII region contains a single, young, hot star; it has broken out of its molecular
cocoon, so it can be seen at optical wavelengths. Compact HII regions have
radii R ∼ 0.5 pc and densities n ∼ 5 × 103 cm−3 . An example of a compact
HII region is the Orion nebula.
A standard HII region contains a single hot star; it typically forms a
‘blister’ on the surface of a molecular cloud. On one side, the expanding
ionized gas of the HII region is confined by the dense molecular cloud; on the
other side, it pushes out into the intercloud medium. Standard HII regions
have R ∼ 3 pc and n ∼ 300 cm−3 . An example of a standard HII region is
the Omega nebula.
A large HII region contains several hot stars; it consists of several stan-
dard HII regions that have merged together. Large HII regions have R ∼
10 pc and n ∼ 50 cm−3 . An example of a large HII region is the North
America nebula.
Finally, a giant HII region contains as many as 104 hot stars; it consists
of numerous large HII regions plus a common ionized envelope. Giant HII
regions have R > ∼ 100 pc and n <
−3
∼ 10 cm . A famous example of a giant HII
region is 30 Doradus in the Large Magellanic Cloud.
Planetary nebulae contain one hot star apiece, but they show an increas-
ing radius and decreasing density similar to that of HII regions. The youngest
visible planetary nebulae have R ∼ 0.03 pc and n ∼ 104 cm−3 . The oldest
planetary nebulae have R ∼ 1 pc and n ∼ 10 cm−3 . The progenitor of a
planetary nebula is a star of initial mass 1 − 8 M⊙ that has evolved onto the
asymptotic giant branch (AGB). At the end of its AGB stage, such a star
64 CHAPTER 6. IONIZATION FRONTS
Basic Turbulence
65
66 CHAPTER 7. BASIC TURBULENCE
A flow that is not turbulent, in which the velocity varies smoothly and
predictably in space and time, is known as a laminar flow. Many flows
can be broken down into a laminar and a turbulent component. Note that
the random turbulent component of a fluid’s motion is distinct from the
random thermal component. The thermal component consists of the random
Brownian motion of the individual gas particles; each particle moves in a
straight line for a distance ∼ λ before an encounter with another particle
sends it off in a different direction. The turbulent component consists of the
eddying motion of macroscopic fluid elements; the eddies have a range of
sizes, but all eddies are very much larger than λ.
By distinguishing between the laminar component and the turbulent com-
ponent, we are separating the bulk velocity ⃗u of a gas into two parts:
⃗ (⃗x) + ⃗u′ (⃗x) .
⃗u(⃗x) = U (7.1)
find a few analytic results! These results will give us some insight into the
physics of the problem, which we can then apply to the more realistic case
of a self-gravitating, radiating, compressible fluid.
In an incompressible fluid, the continuity equation for the turbulent flow
simplifies to the form
∇⃗ · ⃗u′ = 0. (7.4)
The equation for momentum conservation is
∂⃗u′ ⃗ u′ = 1 ∇P
⃗ + ν∇2⃗u′ .
+ (⃗u′ · ∇)⃗ (7.5)
∂t ρ
I am assuming that the fluid exhibits Newtonian viscosity, and that the
kinematic viscosity, ν ≡ µ/ρ, is constant. The kinematic viscosity of air is
ν ≈ 0.15 cm2 s−1 at sea level and at room temperature. For neutral atomic
hydrogen, the kinematic viscosity is
#1/2 " #−1
T n
"
ν ∼ 4 × 1021 cm2 s−1 . (7.6)
104 K 1 cm−3
The kinematic velocity is approximately equal to the sound speed in a fluid
times the mean free path λ.
We can define a velocity correlation tensor for the turbulent velocity
in the following manner:
Since the turbulent velocity field is homogeneous and isotropic (by assump-
tion), the correlation tensor is a function only of the distance r between the
two points and not on their location ⃗x within the velocity field. In an ideal
universe, we could measure the velocity vector everywhere within the tur-
bulent flow, subtract away the laminar component, and then compute the
correlation tensor Rij exactly. Unfortunately, this is difficult to do even in-
side a laboratory. Astronomers are even more badly hampered; usually the
best they can get is a line-of-sight velocity profile, which adds together the
radial velocities of all the matter along a given line of sight.
The trace of the velocity correlation tensor is
$
R(r) ≡ Rii (r) . (7.8)
i
69
In the limit that r → 0, R(r) → ⟨|⃗u′ |2 ⟩. The trace can be used to define a
correlation length
1 !∞
Λt ≡ R(r)dr . (7.9)
R(0) 0
The correlation length, roughly speaking, is the size of the biggest eddies in
the turbulent flow.
Since turbulence exists with a range of eddy sizes, it is frequently conve-
nient to take the Fourier transform of the velocity field in order to consider
the Fourier components of different wavenumber. The components of the
velocity in Fourier space will be
1 ! ′ ⃗
⃗u⃗k = 3
⃗u (⃗r)e−ik·⃗r d3 r . (7.10)
(2π)
The Fourier transform of the velocity correlation tensor is the energy spec-
trum tensor,
⃗ 1 ! ⃗
Φij (k) = 3
Rij (⃗r)e−ik·⃗r d3 r . (7.11)
(2π)
The spectrum Φij (⃗k) tells how much kinetic energy is contained in eddies with
wavenumber k. The tensors Rij and Φij both contain the same information
about the field; which tensor you use depends merely on whether it is more
convenient to work in real space or Fourier space.
In a homogeneous isotropic turbulent flow, it is possible, and also useful,
to define an energy spectrum function E(k) such that
In a Newtonian fluid, the rate at which the turbulent kinetic energy is dissi-
pated by viscosity is ! ∞
ϵd = 2ν k 2 E(k)dk . (7.14)
0
is a pure power law (with α being our old friend, the “factor of order unity”).
Since viscous forces are negligible on this scale, the power spectrum E(k)
must also be independent of the value of the kinematic viscosity ν. Since
1/2 −1/4
u2K = ν 1/2 ϵd and lK = ν 3/4 ϵd , we find
1/2 −1/4 −n/4 n
E(k) = ν 1/2 ϵd ν 3/4 ϵd αν 3n/4 ϵd k . (7.20)
Figure 7.4: The internal velocity dispersion σ plotted versus the maximum
linear size L of molecular clouds. (The different letter symbols represent
different molecular cloud complexes.) [Larson, 1981, MNRAS, 194, 809]
If the relation were actually ∆v ∼ 1 km s−1 (l/1 pc)1/3 , and the molecular
gas were incompressible, that would imply ϵd = (∆v)3 /l ∼ 3×10−4 cm2 s−3 ∼
2 × 10−4 L⊙ / M⊙ . The kinematic viscosity of a dense molecular cloud is
ν ∼ 1017 cm2 s−1 , implying a Kolmogorov length lK ∼ 4 × 1013 cm ∼ 3 AU.
The velocity of the smallest eddies would be uK ∼ 2 × 103 cm s−1 and they
would whirl around on a timescale τK ∼ 2 × 1010 s ∼ 600 yr.
However, treating molecular clouds as if they were incompressible fluids
with a Kolmogorov spectrum of turbulence is incredibly naı̈ve. Most studies
of molecular clouds give the relation
% &0.5
l
∆v ≈ 1 km s−1 (7.26)
1 pc
for the line width on a scale l; the results from one study are shown in
Figure 7.5. In addition, the average density of a molecular cloud scales as
% &−1.2
−3 l
n ≈ 4000 cm . (7.27)
1 pc
Thus, the assumption of a constant density, incompressible fluid is a poor
one. In addition, the turbulent velocities are larger than the sound speed
74 CHAPTER 7. BASIC TURBULENCE
Figure 7.5: Velocity dispersion as a function of linear size for a sample of 273
molecular clouds within our galaxy. [Solomon et al., 1987, ApJ, 319, 730]
75
a ∼ 0.3 km s−1 ; therefore, it is highly likely that shocks will form. Further-
more, energy will be added to the molecular clouds not only on the correla-
tion length Λt , but also on smaller scales (as the result of supernovae, stellar
winds, and other processes). Also, in a realistic view of interstellar turbu-
lence, the effects of gravity and magnetism must be taken into account. In
view of these complications, people who study turbulence in the ISM must
resort to numerical simulations. The simulation shown in Figure 7.6, for
instance, shows the large density contrasts typical of realistic simulations of
compressible turbulence.
76 CHAPTER 7. BASIC TURBULENCE
Figure 7.6: Gas density from a 10243 adaptive mesh simulation of supersonic
compressible turbulence (darker regions are denser). [Kritsuk et al., 2006,
ApJ, 638, L25]
Chapter 8
Spherical Accretion
79
80 CHAPTER 8. SPHERICAL ACCRETION
inward gravitational force on the electron - proton pair is fin = GM∗ mp /r2 .
There exists a limiting luminosity, called the Eddington luminosity, at
which the two forces cancel:
! "
4πGM∗ mp c M∗
LEdd = = 3.4 × 104 L⊙ . (8.2)
σT M⊙
When the luminosity of the central compact object is greater than this value,
the surrounding hydrogen gas will be blown away by the radiation pressure.
If the Eddington luminosity is emitted as black-body radiation, the temper-
ature will be ! "1/4 ! "1/4
LEdd GM∗ mp c
Tbb = = . (8.3)
4πR∗2 σSB R∗2 σSB
where σSB is the Stefan-Boltzmann constant, and R∗ is the radius of the
surface from which the radiation is emitted (for a black hole, of course,
this surface will be outside the Schwarzschild radius). For a black hole ac-
creting at the Eddington limit, the temperature of the radiation will be
Tbb ∼ 4 × 107 K(M∗ / M⊙ )−1/4 , if the radiation comes from immediately out-
side the Schwarzschild radius. The spectrum of the emitted photons will
then peak at a photon energy E ∼ 20 keV(M∗ / M⊙ )−1/4 .
The existence of the Eddington luminosity implies the existence of a max-
imum accretion rate, ṀEdd , for an accreting compact object. If the accre-
tion energy Eacc is converted entirely into radiation, then the luminosity is
Lacc = GM∗ Ṁ /R∗ , and the maximum possible accretion rate is
4πmp cR∗ R∗ R∗
ṀEdd = = 9×1016 g sec−1 ( ) = 1×10−3 M⊙ yr−1 ( ) . (8.4)
σT 1 km R⊙
In reality, the conversion is not 100% efficient, the accretion is not perfectly
spherically symmetrical, and the radiation is not perfectly spherically sym-
metrical; thus, matter can be accreted at rates somewhat greater than ṀEdd .
Quasars and active galactic nuclei can have luminosities as great as 1014 L⊙ .
If they are powered by accretion onto a central compact object, its mass must
be M∗ > ∼ 3 × 10 M⊙ , with a Schwarzschild radius RSch >
9
∼ 8 light-hours. The
X-ray variability of bright AGNs and quasars is on the scale of hours, suggest-
ing that the central accreting object must be a black hole near its Eddington
luminosity; otherwise its radius and variability timescale would be too large.
If the central object is a black hole, the rate at which it must accrete matter in
order to radiate with L = 1014 L⊙ is Ṁ = L/(c2 η) ∼ (η −1 )7 M⊙ yr−1 , where
81
Figure 8.1: Solutions to the Bondi equation, divided into six families.
or
d(u2 ) %%
=0. (8.9)
dr r=rs
%
d(u2 ) %%
Type 3: =0, u2 < a2 . (8.12)
dr r=rs
%
d(u2 ) %%
Type 4: =0, u2 > a2 . (8.13)
dr r=rs
%
83
The type 3 and type 4 families of solution represent flow that is everywhere
subsonic (type 3) or everywhere supersonic (type 4).
u(r)2 a(r)2 GM a2
+ − = ∞ . (8.16)
2 γ−1 r γ−1
a2
! "
1 1
2
a(rs ) + −2 = ∞ , (8.17)
2 γ−1 γ−1
or ! "1/2
2
a(rs ) = a∞ . (8.18)
5 − 3γ
84 CHAPTER 8. SPHERICAL ACCRETION
G 2 M 2 ρ∞
Ṁt = 4πqs , (8.21)
a3∞
where ! "(5−3γ)/(2γ−2)
1 2
qs (γ) = . (8.22)
4 5 − 3γ
The numerical value of qs ranges from qs = 1/4 at γ = 5/3 to qs = e3/2 /4 ≈
1.12 when γ = 1. If the accreting matter is ionized hydrogen, the transonic
accretion rate has the value
! "2 ! "& '−3/2
10 −1 M ρ∞ T∞
Ṁt = 1.2 × 10 g sec . (8.23)
M⊙ 10 g cm−3
−24 104 K
This amounts to about 10−16 M⊙ yr−1 for a 1 M⊙ accreting mass. The Ed-
dington accretion rate for a black hole is ṀEdd ∼ 3 × 1017 g sec−1 (M/ M⊙ ), so
an isolated black hole in the ISM is in little danger of exceeding its Eddington
accretion rate.
The relation between the bulk velocity u(r) and the sound speed a(r) can
be computed from the equation
! "2/(γ−1)
Ṁ Ṁ a∞
−u = = . (8.24)
4πr2 ρ(r) 4πr2 ρ∞ a(r)
Thus "−2/(γ−1)
qs G2 M 2
!
a(r)
u=− 2 3 , (8.25)
r a∞ a∞
85
or "−2/(γ−1)
'−2 !
u qs r a(r)
&
=− , (8.26)
a∞ 4 ra a∞
with ra ≡ 2GM/a2∞ . The “accretion radius” ra is the radius at which the
density ρ and sound speed a start to significantly increase above their ambient
values of ρ∞ and a∞ . The relation between the sonic radius and the accretion
radius is rs = [(5−3γ)/8]ra . At large radii (r ≫ ra ), the infall velocity, sound
speed, and density of the transonic flow are
q s a∞r −2 1 ra
& ' ( )
u ≈ − 1− (8.27)
4 ra 2r
γ − 1 ra
( )
a ≈ a∞ 1 + (8.28)
4 r
1 ra
( )
ρ ≈ ρ∞ 1 + . (8.29)
2r
For a gas with γ = 5/3, the infall velocity, sound speed, and density at small
radii (r ≪ ra ) are
'−1/2
a∞ r
&
u ≈ −a ≈ − (8.30)
2 ra
ρ∞ r −3/2
& '
ρ ≈ . (8.31)
8 ra
If 1 ≤ γ < 5/3, the infall at r ≪ rs is supersonic, and the infalling gas is in
free fall. From the Bernoulli integral, we find that u2 /2 ≈ GM/r, or
'−1/2
r
&
u ≈ −a∞ . (8.32)
ra
The density is then
qs ρ∞ r −3/2
& '
ρ≈ . (8.33)
4 ra
Spherical accretion of gas thus has a characteristic density profile, with ρ ∝
r−3/2 at small radii and ρ = constant at large radii.
The gas will come pouring in with velocity u ∝ r−1/2 at small radii, while
the temperature of the gas increases to the value
'−1
T∞ r
&
T ≈ (8.34)
4 ra
86 CHAPTER 8. SPHERICAL ACCRETION
when 1 ≤ γ < 5/3. The velocity will not increase indefinitely; eventually the
gas will slam into the surface of the compact object – or disappear within
the event horizon if the compact object is a black hole. The velocity of the
gas just before it hits the compact object will be u ∼ −(2GM/R) ∼ −Vesc .
Spherical accretion is only a good approximation for isolated compact
objects. If the accreting mass is fed by a binary companion, or is located in
the center of an active galactic nucleus, an accretion disk will generally form.
The case to which spherical accretion applies is the case of an isolated black
hole or neutron star in the middle of the ISM. It is very unlikely, however,
that a stellar remnant will be at rest with respect to the ambient ISM. Direct
measurements of the radial velocities of pulsars indicate that pulsars in the
disk of our galaxy have a velocity dispersion of ∼ 100 km s−1 . This is much
larger than the sound speed a ∼ 12 km s−1 in the warm neutral medium.
Thus, a typical pulsar will be moving at supersonic speeds with respect to
its ambient medium.
If the accreting body has a velocity V with respect to the ambient medium,
the transonic accretion rate has the form
G 2 M 2 ρ∞
Ṁt = 4π q̃ , (8.36)
(a2∞ + V 2 )3/2
where q̃ is a factor of order unity. When V > a∞ , a bow shock forms in front
of the accreting object, as shown in Figure 8.2, raising the temperature of the
gas and decreasing its bulk velocity relative to the central accreting mass. At
radii r ≪ ra ∼ 2GM/(V 2 + a2∞ ), the flow of the gas is approximately radial,
and takes the form of the spherically symmetric Bondi solution.
A compact object that is moving at highly supersonic speeds will accrete
mass at a rate Ṁ ∝ V −3 . However, there is a lower limit to the accretion
rate that is dictated by the fact that the compact object has a physical cross
section equal to πR2 . Thus, in the limit that the accreting object is moving at
a velocity V that is much greater than the escape velocity Vesc = (2GM/R)1/2
from its surface, the accretion rate is
Figure 8.2: The isovelocity contours (above) and isodensity contours (below)
for accretion onto a compact object moving with speed V = 2.4a∞ with
respect to the ambient gas.
88 CHAPTER 8. SPHERICAL ACCRETION
Figure 8.3: The mass accretion rate for a compact object as a function of
the speed with which is it moving relative to the surrounding gas.
To minimize the rate of accretion for a compact object, you should send it off
at a velocity equal to its escape velocity. A schematic plot of the accretion
rate as a function of the compact object’s velocity is given in Figure 8.3.
Chapter 9
When the gas being accreted has high angular momentum, it generally forms
an accretion disk. If the gas conserves angular momentum but is free to
radiate energy, it will lose energy until it is on a circular orbit of radius
Rc = j 2 /(GM ), where j is the specific angular momentum of the gas, and
M is the mass of the accreting compact object. The gas will only be able to
move inward from this radius if it disposes of part of its angular momentum.
In an accretion disk, angular momentum is transferred by viscous torques
from the inner regions of the disk to the outer regions.
The importance of accretion disks was first realized in the study of binary
stellar systems. Suppose that a compact object of mass Mc and a ‘normal’
star of mass Ms are separated by a distance a. The normal star (a main-
sequence star, a giant, or a supergiant) is the source of the accreted matter,
and the compact object is the body on which the matter accretes. If the
two bodies are on circular orbits about their center of mass, their angular
velocity will be
! "1/2
⃗ = G(Mc + Ms )
Ω ê , (9.1)
a3
∂⃗u ⃗ u = − 1 ∇P
⃗ − 2Ω
⃗ × ⃗u − ∇Φ
⃗ R,
+ (⃗u · ∇)⃗ (9.2)
∂t ρ
89
90 CHAPTER 9. ACCRETION DISKS FOR BEGINNERS
Figure 9.1: Sections in the orbital plane of equipotential surfaces, for a binary
with q = 0.2. The five Lagrange points, L1 through L5 , are shown.
two stars, the equipotential surfaces form a pair of nearly spherical surfaces,
centered on the two stars. It is at intermediate distances that the equipoten-
tial surfaces take on interesting shapes. Of particular interest in the context
of accretion is the ‘critical surface’, which has a dumbbell shape, and consists
of two Roche lobes, one for each star, which are connected at the inner
Lagrangian point L1 . The L1 point is a saddle-point in the potential; the
gravitational acceleration at the L1 point is equal to zero. If the normal star,
in the process of stellar evolution, expands to fill its Roche lobe, then the gas
at the L1 point will be pushed, by its internal gas pressure, from the Roche
lobe of the normal star (Ms ) to the Roche lobe of the compact star (Mc ).
The mass transfer between the two components will change the period of the
binary system. If the total mass and the orbital angular momentum of the
system are constant, then the period of the system is P ∝ Mc−3 Ms−3 . The
period is minimized, given the constraint that Mc + Ms is constant, when
Mc = Ms . Kepler’s law tells us that a ∝ Mc−2 Ms−2 , which is also minimized
when Mc = Ms . When the normal ‘donor’ star is more massive than the
compact accretor (Ms > Mc ), the loss of mass will decrease the separation
a and will decrease the size of the Roche lobe of the normal star; thus, the
mass loss will be naturally self-perpetuating. When Ms < Mc , however, the
separation a will increase as the normal star loses mass, and thus the mass
loss will tend to be cut off (unless evolutionary swelling continues to increase
the size of the mass-losing star, or loss of angular momentum from the binary
system shrinks the separation between the stars.)
The distance of the L1 point from the center of the compact star is well
approximated by the formula
b1
= 0.500 + 0.227 log q , (9.4)
a
so that even when the mass shed by the normal star crosses over the L1
saddle point, it has a considerable distance to travel until it can collide with
the compact star. Moreover, the gas that is passing through the L1 point has
considerable angular momentum. In a nonrotating frame of reference fixed
to the compact star, the L1 point is orbiting with a velocity v ∼ b1 Ω. The
accreting matter that is being squirted through the L1 nozzle therefore has a
bulk velocity u⊥ ∼ v ∼ 2πb1 /P perpendicular to the line between the stars.
Since the accreting matter is being squirted by pressure forces, the velocity of
the gas parallel to the line between the stars will be u∥ ∼ as , where as is the
sound speed in the outermost envelope of the mass-losing star. The normal
92 CHAPTER 9. ACCRETION DISKS FOR BEGINNERS
(The prime denotes differentiation with respect to R). If we use the Keplerian
angular velocity Ω = (GM/R3 )1/2 and the Newtonian viscous torque T =
2πνΣR3 Ω′ , the time evolution of the surface density is
! "
∂Σ 3 ∂ ∂
= R1/2 (νΣR1/2 ) . (9.11)
∂t R ∂R ∂R
This is the basic equation for the evolution of a Keplerian accretion disk.
Once we know the surface density as a function of time, the radial drift
velocity follows from mass conservation:
3 ∂
uR (R, t) = − (νΣR1/2 ) . (9.12)
ΣR1/2 ∂R
Generally, the kinematic viscosity ν will be a function of R and t. How-
ever, the qualitative behavior of a viscous disk can be found by looking at
the simplest case, in which ν = constant. Suppose we start with an accretion
disk that is an infinitesimally thin ring, with mass m and radius R0 :
mδ(R − R0 )
Σ(R, t = 0) = . (9.13)
2πR0
In terms of the dimensionless radius variable x ≡ R/R0 and the dimensionless
time variable τ ≡ t(12ν/R02 ), the solution for the surface density is
1 + x2
! "
m −1 −1/4 2x
% &
Σ(x, τ ) = 2
τ x exp − I1/4 , (9.14)
πR0 τ τ
where I1/4 is a modified Bessel function. This solution is plotted in Figure 9.2
at four increasing values of τ . At early times (τ ≪ 1), the surface density is
nearly Gaussian around R = R0 , with a width σ ≈ (6νt)1/2 . At later times,
most of the mass has lost angular momentum and has moved inward, but
94 CHAPTER 9. ACCRETION DISKS FOR BEGINNERS
Figure 9.2: The viscous evolution of an initially thin ring of mass m and
initial radius R0 . [Pringle, 1981, ARAA, 19, 137]
there is a tail of matter that moves outward to larger and larger radii, and
which carries off most of the angular momentum.
A disk that starts as a thin ring will spread with time. In many cases of
interest, the accretion disk is fed by a steady stream of matter from outside,
and the disk settles into a steady state in which the rate at which gas is lost
from the inner edge is equal to the rate at which gas is added at the outer
edge. In a steady-state disk of this sort, the radial momentum equation is
∂uR u2φ 1 ∂P GM
uR − + + 2 =0, (9.15)
∂R R ρ ∂R R
assuming Keplerian rotation. The inflow velocity uR is of order ν/R, and is
much smaller than the rotation velocity uφ . If we define the Mach number
of the rotation as Mφ ≡ uφ /a,
&1/2
GM
%
uφ = [1 + O(Mφ−2 )] . (9.16)
R
Perpendicular to the disk, the gas is in hydrostatic equilibrium, with
1 ∂P GM z
=− 3 (9.17)
ρ ∂z R
95
or
a2 ∂ρ u2
≈ − φ2 z (9.18)
ρ ∂z R
for a thin disk. Equation (9.18) can be rewritten in the form
1 ∂ρ Mφ2
≈− 2z . (9.19)
ρ ∂z R
This has a solution
z2
! "
ρ ≈ ρ0 exp − 2 , (9.20)
2H
where the scale height is H ≈ R/Mφ . Thus, the thin disk approximation
(H ≪ R) is equivalent to saying that the disk rotation is supersonic (Mφ ≫
1).
In a steady-state disk, the inward mass flux is
Ṁ = −2πRΣuR . (9.21)
Moreover, when ∂Σ/∂t = 0, the equation for the conservation of angular
momentum can be integrated to yield
∂Ω C
−νΣ = −ΣuR Ω − , (9.22)
∂R 2πR3
where C is a constant of integration. For a disk in Keplerian rotation around
a star of radius R∗ , the integration constant is C = Ṁ (GM R∗ )1/2 . Thus,
! &1/2 "
Ṁ R∗
%
νΣ = 1− . (9.23)
3π R
When R ≫ R∗ , the surface density of a Keplerian viscous disk will be Σ ≈
Ṁ /(3πν), and the radial velocity will be uR ≈ −3ν/(2R).
Let D(R) be the rate per unit time per unit area at which the kinetic
energy of rotation is dissipated into heat by viscosity. The value of D is
# $2
1 ∂Ω
D(R) = νΣ R . (9.24)
2 ∂R
For a Keplerian disk, the viscous dissipation rate is
! &1/2 "
3GM Ṁ R∗
%
D(R) = 3
1− . (9.25)
4πR R
96 CHAPTER 9. ACCRETION DISKS FOR BEGINNERS
The formula for the dissipation given above is independent of the value of
ν; it merely includes the assumption that the transfer of mass through the
accretion disk and the dissipation of energy within the disk are both regulated
by the viscosity.
The total disk luminosity is
' ∞ 1 GM Ṁ
Ldisk = 2π D(R)RdR = . (9.26)
R∗ 2 R∗
The disk luminosity is half of the total accretion luminosity Lacc = GM Ṁ /R∗ .
Only half of the luminosity is emitted during the gradual passage of the gas
inward through the accretion disk; the other half must be emitted when the
gas makes the transition from the inner edge of the accretion disk to the
surface of the compact object.
What is the source of the kinematic viscosity ν that causes the inward
drift of the gas in the accretion disk? The first guess would be that ν comes
from standard molecular viscosity, the result of thermal collisions between
individual gas particles in a hot medium. In the case of standard viscosity,
the kinematic viscosity is ν ∼ aT λ, where where aT = (kT /m)1/2 is a typical
thermal velocity, and λ is the mean free path length.
Some more-or-less typical values for accretion disks, at a radius R ∼
1010 cm, are T ∼ 104 K and n ∼ 1016 cm−3 . The mean free path (in an
ionized gas) is
k2T 2
λ = 4 ∼ 10−3 cm , (9.27)
πe n
the thermal velocity in the accretion disk is aT ∼ 106 cm s−1 , and the kine-
matic viscosity is ν ∼ 103 cm2 sec−1 . This value of the kinematic viscosity
yields a viscous accretion time scale of tacc = R2 /ν ∼ 1017 sec ∼ 3 × 109 yr,
and the mass will be flowing in at the excruciatingly slow rate of uR =
−(3ν)/(2R) ∼ −5 cm yr−1 .
To explain the observed accretion rates in X-ray binaries and protostars, a
much larger value of the kinematic viscosity is required. The Reynolds num-
ber for the standard molecular viscosity is Re = uφ R/ν = (GM R)1/2 /ν ∼
1015 at R = 1010 cm from a 1 M⊙ body, with ν = 103 cm2 sec−1 . At such a high
Reynolds number, some physicists have argued, we might expect turbulence
to set in. The random eddies of the turbulence would cause viscosity, just as
random thermal motions cause viscosity on the molecular scale. The kine-
matic viscosity due to turbulence should be, from dimensional arguments,
97
The parameter α is expected, from the arguments given above, to have the
numerical value α < ∼ 1.
The alpha disk is the most commonly used model for accretion disks.
However, by introducing α, we have merely parameterized our ignorance.
There is not even agreement as to the actual source of the viscosity. Some
astronomers, after pointing out that there are no observations that prove that
disks are turbulent, proposed that the viscosity is caused by magnetic stresses
due to a tangled magnetic field within the accretion disk. Such a disordered
field would give rise to an alpha disk with α ∼ vA2 /a2 , where vA is the Alfven
velocity. Other researchers mutter about convective instabilities as a means of
transporting angular momentum; others invoke spiral density waves; others
propose more exotic mechanisms for transporting angular momentum. In
other words, no one knows for sure what causes the viscosity in accretion
disks; we just know that it exists, and that the most plausible mechanisms
result in viscosities that can be written in the form ν = αaH, where α < ∼ 1.
98 CHAPTER 9. ACCRETION DISKS FOR BEGINNERS
Chapter 10
The total luminosity of a disk with the viscous dissipation rate D(R) is
! ∞ 1 GM∗ Ṁ
Ldisk = 2π D(R)RdR = . (10.1)
R∗ 2 R∗
The disk luminosity is half of the total accretion luminosity Lacc = GM∗ Ṁ /R∗ ;
the other half of the luminosity is emitted when the gas makes the transition
from the inner edge of the accretion disk to the surface of the compact object.
The physical characteristics of the boundary layer between the Keplerian
disk and the compact object are poorly known. Let’s start by considering
the better-known luminosity that comes from the disk itself.
If the disk is optically thick, and the luminosity Ldisk is in the form of
black body radiation, then the temperature Tbb (R) at a given radius is set
by the relation
1
σSB Tbb4 = D(R) . (10.2)
2
The factor of 1/2 enters because we are considering the radiation from only
one side of the disk. Using D(R) for a viscous accretion disk,
#−3/4
R
"
Tbb (R) = T∗ [1 − (R∗ /R)1/2 ]1/4 , (10.3)
R∗
where $ %1/4
3GM∗ Ṁ
T∗ = . (10.4)
8πR∗3 σSB
The temperature of the disk is plotted in Figure 10.1. The hottest part of
99
100 CHAPTER 10. ADVANCED ACCRETION DISKS
The spectrum of the disk is the black-body spectrum integrated over all radii:
! Ro ν3
Sν ∝ RdR , (10.6)
R∗ exp[hν/kTbb (R)] − 1
1/3
! hν/kTo x5/3 dx
Sν = ν . (10.7)
0 ex − 1
When kTo /h ≪ ν ≪ kT∗ /h, the upper limit of the integral may be taken as
infinity, yielding the spectrum Sν ∝ ν 1/3 at intermediate frequencies. At very
low frequencies, ν ≪ kTo /h, the main contribution to the spectrum is the
Rayleigh-Jeans tail of the cool, outer edge of the accretion disk; the resulting
frequency dependence is Sν ∝ ν 2 . The integrated spectrum Sν for a disk
with T∗ = 100To is plotted in Figure 10.2.
One check to see whether the ‘physically thin, optically thick’ model is
correct is to check for the characteristic ν 1/3 spectrum. Unfortunately, in
binary systems, the luminosity of the normal star that is providing the ac-
creting matter may overpower the luminosity of the disk. The best system
for seeing the entire spectrum of the accretion disk is a cataclysmic vari-
able in which the compact star is a white dwarf and the normal star is a
low mass (0.1 → 1 M⊙ ) main sequence star. In the quiescent state, between
cataclysmic outbursts, most of the luminosity comes from the disk. The cata-
clysmic variable VW Hydri (a dwarf nova) has a very dim normal companion,
and a large disk. The observed spectrum of VW Hydri, in its post-outburst
state, is shown in Figure 10.3. It is a fairly good example of a ν 1/3 law.
It is enlightening to do slightly more elaborate modeling of alpha disks,
to find its properties as a function of the radius R, the mass M∗ of the central
object, the accretion rate Ṁ , and the assumed value of α. Let ρ, a, P , and
102 CHAPTER 10. ADVANCED ACCRETION DISKS
Figure 10.2: The spectrum Sν of an optically thick accretion disk. The units
of ν and Sν are arbitrary. The temperature at the outer edge of the disk is
Tout = To = 0.01T∗ .
103
Figure 10.3: The dots are the observed spectrum Fλ of VW Hydri after
an outburst. Line ‘a’ is a stellar atmosphere fit; line ‘b’ is a line of slope
Fλ ∝ λ−7/3 , corresponding to Sλ ∝ ν 1/3 .
104 CHAPTER 10. ADVANCED ACCRETION DISKS
3/10 −3/4
T = 1 × 104 K α−1/5 Ṁ16 Mc1/4 R10 f 6/5 (10.20)
1/5
τ = 30 α−4/5 Ṁ16 f 4/5 (10.21)
3/10 3/4
ν = 2 × 1014 cm2 sec−1 α4/5 Ṁ16 Mc−1/4 R10 f 6/5 (10.22)
3/10 −1/4
uR = 3 × 104 cm s−1 α4/5 Ṁ16 Mc−1/4 R10 f −14/5 . (10.23)
Fortunately, none of these results have an exorbitantly strong dependence on
the unknown value of α.
We can check whether our assumption of Kramers’ law is valid. The
opacity within the disk will be
−1/2 3/4
κR (Kramers) = 6.3 cm2 g−1 Ṁ16 Mc−1/4 R10 f −2 . (10.24)
thickness H just outside the boundary layer, which in turn is less than the
radius R∗ of the compact object. To show that b < H < R∗ , start with the
equation for conservation of radial momentum:
In the Keplerian disk, the gravitational term (GM∗ /R2 ) is balanced by the
centrifugal term (u2φ /R). Within the boundary layer, the gravitational term
is balanced by the pressure gradient term. The magnitude of the pressure
gradient is
1 ∂P a2
∼ , (10.29)
ρ ∂R b
where a is the sound speed at the outer edge of the boundary layer and b
is the thickness of the boundary layer. From the balance of forces, a2 /b ∼
GM∗ /(R∗ + b)2 , we see that
R∗ + b
b∼ , (10.30)
Mφ2
107
where Mφ is the rotational Mach number at the outer edge of the boundary
layer. In a thin disk, Mφ ≫ 1, and H ∼ R/Mφ , so b ∼ H/Mφ ∼ R∗ /Mφ2 .
Figure 10.5 shows the resulting geometry, with b ≪ H ≪ R∗ . The black-
body temperature within the boundary layer is
1/4
TBL ≈ Mφ T∗ . (10.31)
Magnetic fields can affect the accretion of matter. The magnetic field of
compact stars is a dipole field, with dipole moment µ = B∗ R∗3 , where R∗ is
the radius of the star and B∗ is the magnetic field at the surface of the star.
A neutron star typically will have R∗ ≈ 106 cm and B∗ ≈ 1012 G, yielding
µ = 1030 G cm3 . A white dwarf with R∗ ≈ 5 × 108 cm and B∗ ≈ 104 G will
have the same magnetic moment, µ ≈ 1030 G cm3 .
At distances r ≫ R∗ , the amplitude of the dipole magnetic field is
The quantity µ30 is the magnetic moment of the star, measured in units of
1030 G cm3 ; the quantity R6 is the radius R∗ of the star, measured in units
of 106 cm.
Consider a neutron star or white dwarf isolated in the middle of a uniform
gaseous medium. The accreting matter will fall in radially at large distances,
where the magnetic pressure is small. The infall will be significantly deflected
from a radial flow at a radius rA (the Alfven radius), where the magnetic
energy density Emag is equal to the kinetic energy density Ekin of the gas.
From the continuity equation,
Ṁ
ρu = − . (10.34)
4πr2
For an ionized gas (γ = 5/3) inside the accretion radius, the infall velocity is
#1/2
GM∗
"
u=− . (10.35)
2r
The kinetic energy density is thus
√
1 2 Ṁ GM∗ −5/2
Ekin = ρu = √ r . (10.36)
2 8π 2
The Alfven radius is %1/7
µ4
$
rA ≈ . (10.37)
Ṁ 2 GM∗
Using the appropriate numerical values for a neutron star or white dwarf,
4/7 −2/7
rA = 7 × 108 cm µ30 Ṁ16 Mc−1/7 . (10.38)
Consider a binary system in which the magnetic field of the compact star
is small enough to allow an accretion disk to form. The accretion disk will
be disrupted at a radius RA , which is the radius at which the torque exerted
by the magnetic field on the disk is equal to the viscous torque. Computing
the magnetic torque is, unfortunately, difficult. The torque exerted depends
on the azimuthal component Bφ of the magnetic field, which in turn depends
on the extent to which the magnetic dipole configuration is distorted by the
interaction of the magnetic field with the disk. Calculations usually show
that RA ≈ rA /2.
Because the magnetic field lines are pinned to the compact object, they
have an angular velocity equal to Ω∗ , the angular velocity with which the
compact object rotates. At radii R ≥ RA , the accreting gas rotates with
an angular velocity Ω(R) = (GM∗ /R3 )1/2 . At radii R ≤ RA , the gas flows
along the magnetic field lines, and hence rotates with an angular velocity
Ω(R) = Ω∗ . For steady accretion to occur, we must have a situation in
3 1/2
which Ω∗ ≤ (GM∗ /RA ) . Numerically, this requires that
−6/7 3/7
Ω∗ < 2 sec−1 µ30 Ṁ16 Mc5/7 . (10.39)
If the magnetized compact object rotates more rapidly than this, gas will be
unable to accrete.
Matter will leave the accretion disk at a radius RA and flow along the
magnetic field lines to the magnetic poles of the neutron star, as shown in
Figure 10.6. The rotation axis of the the neutron star is aligned with the
rotation axis of the disk. However, the magnetic axis of the neutron star is
generally at an angle to the rotation axis of the disk. Let α be the angle
between the magnetic axis of the star and the plane of the disk, as shown in
Figure 10.6. In polar coordinates (r, θ) aligned with the magnetic axis, the
magnetic field lines are described by the equation r(θ) = C sin2 θ. The field
line that passes through the disk at the radius r = RA and angle θ = α must
then have the equation
sin2 θ
r(θ) = RA 2 . (10.40)
sin α
The gas that is ripped away from the disk will follow along this path until
it reaches the surface of the neutron star at a radius r = R∗ . The accretion
thus takes place on a ring of angular radius θc around the magnetic pole,
where θc is given by the relation
R∗
sin2 θc = sin2 α . (10.41)
RA
110 CHAPTER 10. ADVANCED ACCRETION DISKS
Figure 10.6: Accretion from a gaseous disk onto a magnetized neutron star.
111
Diffusion effects tend to smear the area of accretion into a circular cap cen-
tered on the magnetic pole. The area of the two accreting polar caps takes
up a fraction
πR2 sin2 θc R∗ sin2 α
f ∼2 ∗ 2 ∼ (10.42)
4πR∗ 2RA
of the total surface area of the neutron star. The fraction f is small; f ∼ 10−3
for a neutron star with α ∼ π/2. All of the ‘boundary layer’ accretion lumi-
nosity, equal in value to L = (GM∗ Ṁ )/(2R∗ ), will be emitted from two small
regions. The rotation of the neutron star will cause a periodic modulation in
the observed flux from the accreting polecaps. Pulsation periods are observed
in X-ray binaries, with typical periods in the range 1 sec < ∼P <
3
∼ 10 sec. In
many X-ray binaries, the pulsation period P is observed to be decreasing
steadily, on time scales of ∼ 104 yr. This spin-up is the result of the torque
exerted by the accretion disk on the magnetic field of the neutron star. For
a magnetized neutron star accreting matter from a disk, the Keplerian an-
3 1/2
gular velocity at the Alfven radius, Ω = (GM∗ /RA ) , is larger than Ω∗ ,
the angular velocity of the neutron star and its attached magnetic field. The
star + magnetic field combination is thus accreting angular momentum from
2
the disk at the rate Ṁ RA Ω(RA ). If the moment of inertia of the star is
2
I ∼ M∗ R∗ , then the spin-up rate is given by the relation
2
I Ω̇∗ = Ṁ RA Ω(RA ) = Ṁ (GM∗ RA )1/2 . (10.43)
Ṗ 6/7 2/7
− ≈ 10−5 yr−1 I45
−1
Ṁ16 µ30 Mc3/7 P1 , (10.44)
P
where I45 is the moment of inertia in units of 1045 g cm2 .
112 CHAPTER 10. ADVANCED ACCRETION DISKS
Chapter 11
A stellar wind consists of particles emitted from the stellar atmosphere with
a sufficiently large velocity to escape the star’s gravitational attraction. The
escape velocity at the surface of a star with mass M∗ and radius R∗ is
"1/2
2GM∗
!
vesc = = 620 km s−1 (M∗ / M⊙ )1/2 (R∗ / R⊙ )−1/2 . (11.1)
R∗
Since a star will tend to accrete mass, due to its gravitational attraction, some
non-gravitational force is needed to counteract the inward pull of gravity, and
accelerate the outermost layers of the stellar atmosphere away from the star.
This wind-driving force can be, depending on the type of wind, a gradient
in the gas pressure, a gradient in the radiation pressure, or a gradient in the
magnetic pressure.
Observed stellar winds may be placed into one of four main categories.
1. Winds from main sequence stars. Stars similar to the sun have a low
rate of mass loss: Ṁ ∼ 10−14 M⊙ yr−1 . (1 M⊙ yr−1 = 6.3×1025 g sec−1 .)
The asymptotic velocity u∞ of the wind is comparable to the escape
velocity vesc . Winds from solar-type stars are thought to be driven
by gas pressure gradients in the corona. The prototypical star in this
category is the sun.
2. Winds from hot, luminous stars. Stars with Teff >∼ 15,000 K and L > ∼
3000 L⊙ have a high rate of mass loss: Ṁ ∼ 10−7 → 10−5 M⊙ yr−1 .
Their winds are very fast, with u∞ > vesc . Winds from hot stars
are thought to be driven by radiation pressure. Wolf-Rayet stars fall
111
112 CHAPTER 11. THE SOLAR WIND
into this category. WR stars are hot, luminous stars with extended
envelopes; their mass loss rates (Ṁ ∼ 10−5 M⊙ yr−1 ) and wind terminal
velocities (u∞ ∼ 2500 km s−1 ) are among the largest observed for any
type of star.
3. Winds from cool, luminous stars. Stars with Teff < ∼ 6000 K and L > ∼
100 L⊙ have a high rate of mass loss: Ṁ ∼ 10 → 10−5 M⊙ yr−1 . The
−8
wind velocity, however, is very low, with u∞ < vesc . The mechanism
that drives the wind from cool stars is uncertain; a leading candidate
is radiation pressure on dust grains, aided by stellar pulsations in the
outer atmosphere. Stars in this category are K and M giants and
supergiants.
4. Winds from extremely young stars. T Tauri stars have mass loss rates
of Ṁ ∼ 10−9 → 10−7 M⊙ yr−1 . The wind velocity is u∞ ∼ 200 km s−1 .
T Tauri stars have circumstellar disks that might be accretion disks.
The strength of the wind is correlated with the luminosity of the disk;
this suggests that the outgoing wind might be powered by the accretion
disk.
Figure 11.2: The velocity and density of the solar wind at R = 1 AU from
the Sun, as measured by the Mariner 2 spacecraft.
114 CHAPTER 11. THE SOLAR WIND
and the magnetic pressure is B 2 /(8π) = 1.7 × 10−10 ; much smaller than the
kinetic energy density ρu2 /2 = 9.4 × 10−9 erg cm−3 .
At any given time, part of the sun’s corona is emitting a low-speed wind,
and part is emitting a high-speed wind. (The high-speed winds appear to
come from “coronal holes” – regions of low density and low temperature
where the magnetic field lines are not closed.) Both the high and low speed
winds produce a proton flux of ∼ 3 × 108 protons cm−2 sec−1 at 1 AU. This
leads to a total mass loss rate of 2 × 10−14 M⊙ yr−1 .
The study of stellar atmospheres usually starts with the statement, “We
assume that the atmosphere is in hydrostatic equilibrium.” Well, that state-
ment is fraudulent – a stellar atmosphere cannot be in hydrostatic equilib-
rium. There is always going to be a loss of high velocity particles. To show
that the sun’s corona is not in hydrostatic equilibrium, we first assume that
it is in equilibrium, and then show that the assumption leads to an absurd
conclusion. The corona is the outermost part of the sun’s atmosphere; it
extends from a distance of 2000 km above the sun’s visible surface out into
interplanetary space. At the base of the corona, r0 = 7×1010 cm, the temper-
ature is T0 = 2 × 106 K and the number density of protons is n0 = 108 cm−3 .
The resulting gas pressure is P0 = 2n0 kT0 = 0.06 dyne cm−2 . At the high
temperatures present in the corona, the thermal structure is determined by
heat conduction. The coefficient of thermal conductivity for an ionized gas
is equal to
"5/2
T
!
9 −3 −1
K(T ) = 3 × 10 g cm sec K . (11.2)
2 × 106 K
In the corona, where there are no local heat sources, the fact that heat flow
is in a steady state tells us that
# $
∇ ⃗ ) = 1 d r2 K dT
⃗ · (K ∇T =0. (11.3)
r2 dr dr
Figure 11.3: The wind velocity of an isothermal Parker wind, for different
values of the temperature T0 .
u2 u2
# $
r rs
− ln = 4 ln +4 −3 . (11.17)
a20 a20 rs r
⃗ ·B
B ≈ 5 × 10−5 G. The fact that ∇ ⃗ = 0 implies that
⃗ = Br (r)êr + Bφ (r)êφ
B (11.20)
⃗ = −⃗u × B/c,
The electric field in the ionized wind is E ⃗ and the steady-state
Maxwell’s equation ∇ ⃗ ×E ⃗ = 0 reduces to the equation
1d
[r(ur Bφ − uφ Br )] = 0 . (11.22)
r dr
119
Integrating, we find
d Br r2 d
(ruφ ) = (rBφ ) . (11.24)
dr Ṁ dr
Since Br r2 = constant, we may instantly perform the integration to find that
r2 Br
# $
r uφ − Bφ = L = constant . (11.25)
Ṁ
The constant L is the specific angular momentum carried away by the solar
wind. The first term in the above equation is the specific angular momentum
carried by the motions of the gas; the second term represents the angular
momentum carried by the magnetic field.
It is convenient to define the radial Alfvenic Mach number, MA ≡ ur /vA ,
where vA is the Alfven velocity,
$1/2
Br2
#
vA = . (11.26)
4πρ
At the base of the corona, vA ≈ 300 km s−1 ; at 1 AU, the Alfven velocity has
fallen to vA ≈ 40 km s−1 . The azimuthal velocity of the solar wind, in terms
of Ω∗ , MA , and L, is
At the base of the corona, where the radial velocity ur is still small, the
Alfvenic Mach number is much smaller than one. At 1 AU, the Alfvenic
Mach number is MA ≈ 10. Thus, somewhere between the sun and the earth,
there exists an Alfvenic critical radius at which MA = 1. Let the radius at
this point be rA and the radial velocity be uA . Since the denominator in
the equation for uφ vanishes at rA , the numerator must vanish, too. Thus,
2
L = Ω∗ rA . The angular momentum per unit mass within the solar wind can
120 CHAPTER 11. THE SOLAR WIND
J˙ = −LṀ = −Ω∗ rA
2
Ṁ . (11.28)
For typical models of the magnetized solar wind, the Alfven radius is rA =
1.7 × 1012 cm = 24 R⊙ = 0.11 AU. With the sun’s angular velocity of Ω∗ =
3 × 10−6 sec−1 and mass loss rate of Ṁ = 2 × 10−14 M⊙ yr−1 , this implies
Since the sun’s total angular momentum is J = 1.6 × 1048 g cm2 sec−1 , the
time scale over which the sun will be spun down is
This is a time comparable to the age of the sun, in contrast to the mass loss
time scale tM = M⊙ /Ṁ ≈ 5 × 1013 yr. Although the solar wind will not
significantly affect the total mass of the sun, it will affect the total angular
momentum of the sun.
The angular velocity of the wind may be rewritten as
Ω∗ r uA − ur
uφ = (11.31)
uA 1 − MA2
Figure 11.4: The angular momentum content of the gas in the solar wind
(solid line) compared to the angular momentum content of the magnetic field
(dashed line).
the vicinity of the Alfven radius. The azimuthal component of the magnetic
field decreases at the rate Bφ ∝ r−1 . Since Br ∝ r−2 , at large radii, the
magnetic field will be mainly azimuthal; at small radii, the field will be
mainly radial.
At a radius r, the gas in the solar wind has a specific angular momentum
Ω∗ r2 uA − ur
Lgas = uφ r = , (11.33)
uA 1 − MA2
plotted as the solid line in Figure 11.4. The magnetic field carries the specific
angular momentum
2
rA − r2
Lmag = L − Lgas = Ω∗ , (11.34)
1 − MA2
which is shown as the dashed line in Figure 11.4. The ratio of angular
momentum in the gas to angular momentum in the magnetic field is, at
small radii, # $2
! "2
r −3 r
Lgas /Lmag ≈ ∼ 2 × 10 . (11.35)
rA R⊙
122 CHAPTER 11. THE SOLAR WIND
Figure 11.5: The termination shock, heliopause, and bow shock surrounding
the solar sytem.
if the pressure of the ISM is taken to be Pi ≈ 3×10−13 dyne cm−2 . The space-
craft Voyager 1 went through this termination shock in 2004 December,
when it was 94 AU from the Sun (Figure 11.5). Voyager 2 went through the
termination shock in 2006 May, when it was only 76 AU from the Sun; this
gives a idea of the non-sphericity of the termination shock.
Chapter 12
Hot luminous stars, such as O and B supergiants, are known to have stellar
winds. In the UV, they display P Cygni line profiles.1 P Cygni profiles show
absorption at short wavelengths and emission at longer wavelengths, as seen
in the lower panel of Figure 12.1. This asymmetric absorption/emission
profile is the characteristic signature of an expanding stellar atmosphere.
The absorption comes from material between us and the photosphere of the
star (region 5 in the upper panel of Figure 12.1). The emission comes from
material in regions 1, 2, 3, and 4. The width of the P Cygni line tells us
the terminal velocity u∞ of the expanding gas. The depth of the absorption
component tells us the column density of absorbing matter along the line of
sight to the star. From this, we can use a model of the expanding atmosphere
to deduce the total mass of the gas surrounding the star. The P Cygni lines
of OB supergiants show expansion velocities of ∼ 2000 km s−1 , and mass
loss rates of ∼ 10−6 M⊙ yr−1 . The hottest stars can have maximum wind
velocities of as much as 4000 km s−1 .
Mass loss from hot luminous stars cannot be explained by gas pressure
gradients like those that drive the solar wind. Among the P Cygni lines that
are observed in OB supergiants are those of CIV and SiIV. These ions are
present at a temperature of ∼ 3 × 105 K; if the gas were any hotter, they
would be collisionally ionized to higher ionization states. In a hot coronal
wind driven by gas pressure, the maximum observed velocity can be only a
few times the sound speed, which is a ∼ 70 km s−1 at T ∼ 3 × 105 K. Thus,
1
P Cygni itself is a variable blue hypergiant (B1 Ia); it is visible to the naked eye,
despite being some 2 kpc from the Earth. It’s an example of the type of hot, luminous
star that produces a fast stellar wind with a high mass-loss rate.
123
124 CHAPTER 12. WINDS FROM HOT & COOL STARS
we need some additional force to explain the very high wind velocities seen
in OB supergiants.
Winds in hot luminous stars are driven by radiation pressure. The stellar
winds emerging from OB supergiants have numerous resonance lines in the
UV, which coincidentally is where the continuum radiation of an OB star has
its maximum. The winds from OB stars are thus referred to as line-driven
winds, since the opacity of the accelerated material is provided by absorption
lines. A rough estimate of the mass-loss rate in a radiatively driven wind can
be computed by assuming that each photon emitted by the star transfers
its momentum of hν/c to a gas particle in the wind. The rate at which the
star loses momentum is L/c, where L is the radiative luminosity. The rate
at which the wind carries away the momentum is Ṁ u∞ , where u∞ is the
asymptotic wind velocity. By setting the two rates equal to each other, we
estimate a mass loss of
L
Ṁest = (12.1)
u∞ c
and a corresponding kinetic energy luminosity of
1 u∞
Ṁest u2∞ = L. (12.2)
2 2c
For a luminous O or B star, this estimated mass loss rate is Ṁest ∼ 10−6 →
10−5 M⊙ yr−1 . This is in the right range to account for the observed mass loss
from OB stars. The kinetic energy luminosity is less than 1% of the radiation
luminosity L. If each photon from the star undergoes multiple scatterings,
then the mass loss rate can be several times Ṁest .
Even if the circumstellar gas had no resonance lines, there would still be
radiation pressure on the ionized gas as a result of Thomson scattering. A
star of luminosity L will exert an outward radial force of magnitude f =
(σT L)/(4πr2 c) on every electron-proton pair (where σT = 6.7 × 10−25 cm2 is
the Thomson cross section). The ratio of the outward acceleration due to
Thomson scattering, to the inward acceleration due to gravitation, is
σT L 3.2 × 10−5
Γ= = , (12.3)
4πGM mp c M/L
drive the stellar winds. For a very luminous O star, with L ≈ 106 L⊙ and
M ≈ 60 M⊙ , you find Γ ≈ 1/2; for less luminous stars, Γ will be even smaller.
A single line will provide an acceleration
κL ! ∞
gL = πFν (ν)φ(ν)dν , (12.4)
c ν=0
where κL is the line opacity (in units of cm2 g−1 ), πFν is the flux from the
central star at frequency ν (in units of erg sec−1 cm−2 Hz−1 ) and φ(ν) is the
line profile function.
A complication is added by the fact that the gas of the stellar wind
is being steadily accelerated away from the star. Thus, the radiation that
each bit of gas sees from the central star will be more and more redshifted.
Suppose that the gas has a strong line at a frequency ν0 , as measured in the
rest frame. A photon emitted by the star with a frequency ν > ν0 will be
scattered by the shell of gas that has a velocity
ν − ν0
u(r) = c . (12.5)
ν0
The stellar wind will be able to scatter photons with frequencies between
ν0 and ν0 + δν, where δν = ν0 (u∞ /c). Photons with a frequency ν0 will be
scattered near the surface of the star; photons with a frequency ν0 + δν will
be scattered far from the star, where the expansion velocity has reached its
asymptotic value of u∞ .
That maximum mass loss rate that can be produced by a single strong
line is found by setting the mass momentum flux equal to the radiation
momentum flux in the frequency range ν0 → ν0 + δν:
Lν δν L ν ν0
Ṁ u∞ = = 2 u∞ . (12.6)
c c
The quantity Lν δν is the stellar luminosity in the frequency range ν → ν +δν.
If ν0 is near the peak of the star’s energy distribution, then Ṁ ∼ L/c2 . This
rate is smaller by a factor u∞ /c than the mass loss rate Ṁest that occurs
if all photons are scattered or absorbed, rather than those in the restricted
frequency range ν0 → ν0 (1 + u∞ /c). Even for the fastest stellar winds, the
ratio u∞ /c is only 0.01. If a line-driven wind is to effectively convert the
momentum of photons into the momentum of gas, it must have many strong
lines in the frequency range where the star emits most of its radiation.
127
Suppose that a radiative flux Fν is incident on the inner side of a thin gas
shell. If the gas scatters photons with a frequency ν0 , then the observed res-
onance line will have a Doppler width ∆νD = ν0 ut /c, where ut = (kT /m)1/2
is the thermal velocity within the shell. The radiative acceleration of the
shell is
πFν 1 − e−τL
gL = κL ∆νD . (12.7)
c τL
In the above equation, τL is the effective optical depth of the shell, and is
given approximately by the equation
" #−1
du
τL = κL ρ(r)ut (r) . (12.8)
dr
πFν ν0 du
gL ≈ , (12.9)
c ρc dr
πFν ν0 κL ut
gL ≈ . (12.10)
c c
In real stellar winds, the total line acceleration,
$
gL = gL,i , (12.11)
i
will be a sum over weak and strong lines. It is customary (following Cas-
tor, Abbott, & Klein 1975, ApJ, 195, 157), to parameterize the total line
acceleration in a spherically symmetric wind as
" #α
GM 1 du
% &
gL = 2
Γ k , (12.12)
r κρut dr
where the force constant k and the slope α are found by an empirical fit to the
observed line strengths in stellar atmospheres. (The parameter κ = σT /mp =
0.40 cm2 g−1 is the Thomson scattering opacity.) If all the observed lines are
strong, α = 1; if all the lines are weak, α = 0. An O4 star has k ≈ 1/30 and
α ≈ 0.7.
128 CHAPTER 12. WINDS FROM HOT & COOL STARS
With this parametric form for the line acceleration, the isothermal Bondi
equation takes the modified form
1 d(u2 ) a2 2a2 r
' (
GM
(1 − 02 ) = − 2 1 − 0 − Γ − Γkt−α , (12.13)
2 dr u r GM
where #−1
r2 d(u2 )
" #"
κut Ṁ
t= . (12.14)
4π 2 dr
In the absence of line opacity, the solution to this equation is a Parker wind
with a sonic radius at
GM (1 − Γ)
rs = . (12.15)
2a20
Once the line opacity is added, the sonic radius is no longer a critical point.
There is, however, a unique solution which has a smooth transition from
subsonic flow near the star to supersonic flow far from the star. The terminal
velocity for this model has the value
&1/2
α
%
u∞ = vesc , (12.16)
1−α
where ' (1/2
2GM (1 − Γ)
vesc = . (12.17)
R∗
The observed values of u∞ /vesc for a sample of hot, luminous stars are plotted
in Figure 12.2. The O stars are observed to have u∞ /vesc ≈ 3, suggesting a
value of α ≈ 0.9 for these stars. Late B stars, by contrast, have u∞ /vesc ≈ 1,
yielding α ≈ 0.5. The mass loss rate for a line-driven wind is
&(1−α)/α
L 1/α Γ
%
Ṁ = k α(1 − α)(1−α)/α . (12.18)
ut c 1−Γ
This leads to a luminosity dependence Ṁ ∝ L1/α . If all the lines are optically
thick, then Ṁ ∝ L; if all the lines are optically thin, then there will be a
very steep dependence of Ṁ on L.
An O5 star has a mass M = 60 M⊙ , a luminosity L = 106 L⊙ , and a radius
R∗ = 14 R⊙ . The effective temperature is then 49,000 K. Castor, Abbott,
and Klein used their parametric model, with k = 1/30 and α = 0.7, to deduce
129
Figure 12.2: The ratio of terminal wind velocity to escape speed as a function
of spectral type for stars from O4 to A2.
130 CHAPTER 12. WINDS FROM HOT & COOL STARS
a terminal velocity u∞ = 1500 km s−1 for such a star, and a mass loss rate
of Ṁ = 7 × 10−6 M⊙ yr−1 . This is in good agreement with observations of O
stars. The time scale for mass loss in such a star is tM = M/Ṁ = 1 × 107 yr,
which is only three times the main sequence lifetime of an O5 star.
The terminal velocities deduced for cool luminous stars range from u∞ ∼
10 km s−1 for M supergiants to u∞ ∼ 75 km s−1 for K giants. These velocities
are smaller than the corresponding escape velocities vesc from the stellar
surface. An approximate fit, found empirically, is
2
vesc
u∞ ∼ . (12.19)
1000 km s−1
The deduced mass loss rate for cool luminous stars lies in the approximate
range 10−8 M⊙ yr−1 → 10−5 M⊙ yr−1 .
Winds in cool K and M stars cannot be driven by pressure gradients, since
cool stars with winds lack hot, extended coronas. Thomson scattering is not
enough to drive the winds of cool stars. An M supergiant, with L = 3×104 L⊙
and M = 20 M⊙ , will have Γ = L/LEdd = 3.2 × 10−5 (M/L)−1 = 0.05. A K
giant, with L = 130 L⊙ and M = 4 M⊙ , will have Γ = 0.001. The radiative
acceleration from lines is also small; the continuum of cool stars peaks in
infrared or red, where there are few resonance lines in the atmosphere. How-
ever, in cool giants and supergiants, there is an additional source of opacity.
Cool giants and supergiants are observed to emit an excess of radiation in
the infrared. This infrared excess can be attributed to the formation of
dust grains in the cool, extended stellar envelope; dust condensation requires
temperatures lower than T ∼ 1000 K.
The infrared opacity of the atmosphere increases very rapidly at the radius
where the dust condenses out. The grains, once they form, are accelerated
outward as the result of the momentum they gain by absorbing photons.
Some of the momentum of the grains is transferred to the gas by collisions.
As the grains are driven outward, they thus drag the gas along with them.
When the dust grains are accelerated to high velocities, however, they begin
to be destroyed by ‘sputtering’. The acceleration ceases, and the terminal
wind velocity stays at the relatively low value of ∼ 20 − 50 km s−1 . If dust
particles were indestructible, the wind could be accelerated to velocities of
∼ 100 km s−1 or more. The radiative acceleration on the dust grains is greater
than the gravitational acceleration when
4πGM c M
% &
κd > = 1.3 × 104 cm2 g−1 , (12.20)
L L
131
Figure 12.3: The density (upper panel) and velocity (lower panel) for shock-
driven winds. On the left, the wind is assumed to be adiabatic; on the right,
the wind is assumed to be isothermal. [Wood 1979, ApJ, 227, 220]
more momentum and energy is added to the gas. Eventually, the outer layers
of the star are unbound, and flow outward in a wind. The properties of a
shock-driven wind, as illustrated in Figure 12.3, depend on whether the gas
is adiabatic or isothermal. A model star with M = M⊙ , L = 104 L⊙ and pul-
sational period P = 373days will have a mass loss rate of Ṁ = 0.02 M⊙ yr−1
if the gas is adiabatic, and a mass loss rate of Ṁ = 10−12 M⊙ yr−1 if the gas is
isothermal. The main difference in the two types of flow is that the adiabatic
case has much higher densities at large radii. A combination of isothermal
shock-driven flow (effective at small radii) and dust-driven flow (effective at
large radii) gives a mass loss rate of Ṁ = 3 × 10−7 and a terminal velocity
u∞ = 7 km s−1 , which are of the right magnitude for mass loss from a Mira
variable.
Chapter 13
Astrophysical Jets
1. Young stellar jets are seen near newly forming stars. The length of a
young stellar jet is l = 0.01 → 1 pc, and the velocity of gas within the
jet is v = 100 → 400 km s−1 ; the central compact object is a newborn
star of mass M ∼ 1 M⊙ . A young stellar jet carries off mass at a rate
Ṁ = 10−9 → 10−6 M⊙ yr−1 .
2. Extragalactic jets are seen near the nuclei of radio galaxies and
quasars. The length of an extragalactic jet is l = 0.01 → 1 Mpc, and
the velocity of gas within the jet may be as high as v ∼ c; the central
compact object is (probably) a black hole of mass M = 108 → 109 M⊙ .
Despite the very different length scales and velocities for the two types
of jets, the basic physics involved is the same. Long, highly collimated flows
originate in a compact object, and appear to be perpendicular to an ac-
cretion disk. In addition to a force that counteracts the gravitational force
of the central compact object, the formation of a jet requires a ‘nozzle’ to
shape the gas flow into a narrow jet. Theorists have divided the study of
astrophysical jets into two parts. (1) What forces act to accelerate and col-
limate the beam? (2) How does the jet propagate through space once it has
been accelerated and collimated? The first question is the more difficult one.
Jets are presumably accelerated by one of the mechanisms that accelerate
133
134 CHAPTER 13. ASTROPHYSICAL JETS
Figure 13.1: The components of a typical jet from a young stellar object.
of the jet, the ambient gas is moving with a velocity −uw and the jet gas
is moving with a velocity uj − uw . Since the working surface is stationary
in this frame, the ram pressure of the gas on the two sides of the working
surface must be equal. That is,
η 1/2
uw = uj . (13.3)
1 + η 1/2
For heavy jets, with η ≫ 1, the working surface moves outward with a
velocity uw ≈ uj . For light jets, with η ≪ 1, the working surface moves
outward with the low velocity uw ≈ uj η 1/2 . Observations of the velocities of
young stellar jets, and of the Herbig-Haro objects at their working surfaces,
136 CHAPTER 13. ASTROPHYSICAL JETS
indicate that η ∼ 1 − 2. The density of hydrogen atoms in the stellar jet and
in the ambient molecular cloud are n ∼ 50 cm−3 .
The age of a young stellar jet, as deduced from its expansion velocity, is
! "# $−1
3 R uw
tj ≈ 10 yr , (13.4)
0.1 pc 100 km s−1
where R is the length of the jet. The mass of the jet is
! "# $−1 ! "
R uw Ṁ
Mj ≈ 10−4 M⊙ , (13.5)
0.1 pc 100 km s−1 10 M⊙ yr−1
−7
In addition to the dependence on the density ratio η, the structure of the jet
also depends on its Mach number Nj = uj /aj , where aj = (γPj /ρj )1/2 s. Jets
with low Mach numbers have cocoons that extend only a short distance from
the working surface; jets with high Mach numbers have cocoons that extend
for the entire length of the jet. Figure 13.2 shows a series of jet simulations;
all the jets have Mach number mj = 6, but their value of η ranges from
η = 10 (no cocoon worth mentioning) to η = 0.01 (a cocoon much larger
than the central jet). Figure 13.3 shows a series of jet simulations in which
all the jets have η = 0.1, but their Mach number ranges from mj = 1.5 (short
cocoon) to mj = 12 (cocoon as long as the jet).
Since the jet has an internal pressure Pj , what keeps the jet from expand-
ing perpendicular to its long axis? In non-magnetic jets, the confinement is
provided by the external pressure Pex . If the pressure of the external medium
is constant, then the diameter d of the jet is roughly constant as a function of
the distance r along the jet. If, however, the external pressure Pex varies with
distance along the jet, the diameter of the jet will vary in order to maintain
pressure equilibrium at the sides of the jet. Suppose that the gas within the
jet is polytropic, with
Pj = P0 (ρj /ρ0 )γ . (13.11)
If the external pressure has a power-law dependence on r,
If the mass flux Ṁ and the bulk velocity uj are constant along the jet, then
the mass continuity relation
π
Ṁ = d(r)2 ρj (r)uj (13.14)
4
tells us that ! "1/2 # $n/(2γ)
4Ṁ r
d(r) = . (13.15)
πuj ρ0 r0
The opening angle θ of the jet is θ ≈ d/r as long as d ≪ r. When the jet is
narrow,
# $(n−2γ)/(2γ)
r
θ(r) = θ0 . (13.16)
r0
138 CHAPTER 13. ASTROPHYSICAL JETS
When γ > 1, the Mach number of the jet increases as the external pressure
decreases.
If the jet has a very high Mach number, or is very broad, it can no longer
be confined by the pressure of the ambient medium. Place yourself in a
frame of reference traveling along the axis of symmetry of the jet with the
gas velocity uj . In this frame of reference, the walls of the jet are moving
away from you with a velocity
! "
θ θ
uwall ≈ uj tan ≈ uj . (13.18)
2 2
Information about changes in the external pressure are carried toward you
from the wall of the jet with a velocity equal to the sound speed aj . In
order for the jet to remain confined by the pressure of the outside gas, it
must remain in causal contact with the gas outside the walls of the jet. This
requires that uwall < aj , or, in other words, that θmj < 2. In our model with
Pex ∝ r−n , the product of the opening angle θ and the Mach number mj is
$(n−2)/2
r
#
θmj = θ0 m0 . (13.19)
r0
Thus, when n > 2, there will be a critical radius rf at which θmj = 2; beyond
this point, the jet will no longer be confined by the pressure of the ambient
medium. It will be a free jet. Free jets have a constant opening angle
2
θf = . (13.20)
mj (rf )
The diameter of the free jet will increase at the rate d ∝ r, the density will
decrease at the rate ρj ∝ r−2 , and the Mach number will increase at the rate
mj ∝ rγ−1 . Some extragalactic jets have a constant opening angle, and are
conjectured to be free jets. One such jet is seen in NGC 6251 (Figure 13.4).
Radiative cooling can have an important effect on the structure of stellar
jets. Cooling creates a dense, cool shell at the head of the jet. The density
within the shell, simulations indicate, can be as much as 100 times the density
141
Figure 13.4: The jet of the radio galaxy NGC 6251, seen at radio frequencies
at over a wide range of angular resolutions.
142 CHAPTER 13. ASTROPHYSICAL JETS
of the ambient medium. This radiating, dense shell can be identified with
Herbig-Haro objects seen at the ends of young stellar jets. Cooling also
decreases the size and the pressure of the cocoon.
Jets are not perfectly smooth and straight; they show a tendency to
break up into knots and to deviate into ‘wiggles’. These knots and wiggles
can occur because the jet is subject to instabilities that are analogous to the
classic Kelvin-Helmholtz instability. Figure 13.5 shows the development of
jet wiggles in a numerical simulation. Numerical studies indicate that there
exist growing modes of instability for jets with any Mach number. Some
of these modes are axisymmetric, pinching modes (these would tend to give
rise to observed knots) and some modes are nonaxisymmetric helical modes
(these would tend to give rise to observed wiggles). The fastest growing
modes are those with wavelengths comparable to the diameter d of the jet
(this is why knots are roughly circular and wiggles have radii of curvature
roughly equal to the jet width).