Mit8 902 f23 Lec Full
Mit8 902 f23 Lec Full
902: Astrophysics II
Massachusetts Institute of Technology
Department of Physics
Lecture notes by
Mark Vogelsberger, Stephanie O’Neil, & David DePalma
Fall 2023
Contents
I Galaxies 3
1 Key observations of galaxies . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.A Basic units of radiative transfer . . . . . . . . . . . . . . . . . . . . . 4
1.B Basic properties of the galaxy population . . . . . . . . . . . . . . . . 9
1.C Stellar population synthesis . . . . . . . . . . . . . . . . . . . . . . . 11
2 Structure and a qualitative picture of galaxies . . . . . . . . . . . . . . . . . 12
2.A Virial Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.B Relaxation times . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.C Collisionless relaxation . . . . . . . . . . . . . . . . . . . . . . . . . . 19
3 Modelling galaxies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
3.A Potential-density pairs . . . . . . . . . . . . . . . . . . . . . . . . . . 22
3.B Orbits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.C Phase-space distribution function . . . . . . . . . . . . . . . . . . . . 37
3.D Stability of stellar systems . . . . . . . . . . . . . . . . . . . . . . . . 42
3.E Stellar population synthesis . . . . . . . . . . . . . . . . . . . . . . . 48
3.F Chemical evolution of galaxies . . . . . . . . . . . . . . . . . . . . . . 52
3.G Active galaxies (AGN) . . . . . . . . . . . . . . . . . . . . . . . . . . 56
1
CONTENTS CONTENTS
2
Part I
Galaxies
3
1. KEY OBSERVATIONS OF GALAXIES
Flux:
dEν
Fν = (1)
dA dt dν
with units [Fν ] = erg s−1 cm−2 Hz−1 .
Fν is the flux at a specific frequency ν.
Specific intensity:
dEν
Iν = (2)
dA dt dν dΩ
with units [Iν ] = erg s−1 cm−2 Hz−1 sr−1 .
This is the flux per solid angle.
Note:
• Fν (r) ∝ r12 .
Due to energy conservation:
Fν (r1 ) × 4πr12 =
(dEν )1 = (dEν )2 = (3)
Fν (r2 )2 × 4πr22
2
Fν (r1 ) r2
⇒ =
Fν (r2 ) r1 (4)
1
⇒ Fν (r) ∝ 2
r
• Iν (r) ∝ constant because:
Fν 1
Iν = and dΩ ∝ 2
dΩ r
1 1 (5)
⇒ Fν ∝ 2 and dΩ ∝ 2
r r
⇒ Iν ∝ constant
4
1. KEY OBSERVATIONS OF GALAXIES
Magnitude scale:
Define the apparent magnitude m, i.e. how bright an object appears:
(Fν )1
m1 − m2 = −2.5 log (6)
(Fν )2
With this definition, a brighter object has a lower magnitude. There are two main magnitude
systems: Vega and AB.
The Vega system is calibrated using the flux of the AO V star Vega (Fν )Vega , which has a
non-flat distribution (flux changes for different frequencies). The AB system is calibrated to
We also have the monochromatic magnitude, i.e. the magnitude at a single wavelength,
defined for each system:
Fν
Vega : mν = −2.5 log
(Fν )Vega
(8)
Fν
AB : mν = −2.5 log
(Fν )AB
R
Fν TX (ν)dν
Vega : mX = −2.5 log R
(Fν )Vega TX (ν)dν
R (9)
Fν TX (ν)dν
AB : mX = −2.5 log R
(Fν )AB TX (ν)dν
R
(Fν )AB = constant and TX (ν)dν = 1, so
Z
(Fν )AB TX (ν)dν = (Fν )AB . (10)
5
1. KEY OBSERVATIONS OF GALAXIES
Telescopes like Hubble and SDSS observe primarily in the visible light spectrum. JWST
measures slightly longer wavelengths and is sensitive to the infrared range. The NIRCam
instrument filters are shown in below. We show the total throughput (photon-to-election
conversion efficiency) for extra-wide, wide, medium, and narrow filters for NIRCam (image
from https://jwst-docs.stsci.edu).
Each filter measures a different energy range of electromagnetic waves and therefore probes
different physics. As an example of this, we show the Orion Nebula as viewed in visible light
from Hubble below on the left and in X-ray from Chandra on the right. In the visible range,
we can see the diffuse gas while in the X-ray, we can see point-like sources from stars.
Figure is in the public domain. JWST User Documentation (JDox). Baltimore, MD: Space Telescope
Science Institute; 2016-2024-07-25. https://jwst-docs.stsci.edu
6
1. KEY OBSERVATIONS OF GALAXIES
For all filters, −2.5 log(3.63 × 10−20 ) = 48.6, so (for the AB system)
Z
mX = −2.5 log Fν TX (ν)dν − 48.6
(11)
mν = −2.5 log(Fν ) − 48.6 .
The value for (Fν )AB was chosen such that mV (AB) = mV (Vega) and they have the same
magnitude in the V-band. For other bands, one must apply the conversion
R
(Fν )AB TX (ν)dν
mX (AB) − mX (Vega) = −2.5 log . (12)
(Fν )Vega TX (ν)dν
Define the absolute magnitude as the apparent magnitude if the object were at a distance of
10 pc. Apparent magnitude depends on both the brightness of the object and its distance.
Absolute magnitude is related to the intrinsic brightness of the object.
D
mX − MX = 5 log ≡µ (14)
10 pc
Image of Orion Nebula created by Mark Vogelsberger using SAOImage DS9 image display and
visualization tool for astronomical data. https://sites.google.com/cfa.harvard.edu/saoimageds9/home
7
1. KEY OBSERVATIONS OF GALAXIES
Since
L L
Fapp = and
Fabs =
4πD2 4π(10 pc)2
L
⇒ mX = −2.5 log + constant
4πD2
L (15)
MX = −2.5 log + constant
4π(10 pc)2
⇒ mX − MX = +2.5 log(D2 ) − 2.5 log((10 pc)2 )
= 5 log(D) − 5 log(10 pc)
= 5 log(D/pc) − 5
µ = mX − MX is the distance modulus (a measure of distance).
Colors:
If observations are made in more than one filter (X, Y ), then one can define a color as the
difference in magnitudes between the two bands:
(X − Y ) = mX − mY = MX − MY (16)
Stars and galaxies can be “red" or “blue", for example. It is common to use the difference
between g and r filters to get g − r color. A higher g − r value is red and a lower value
is blue. Note that higher g − r has a higher g relative to r, but a higher magnitude is less
bright.
We can take images of the same object through different filters and combine them for a more
complete view of the object. Here we show images of the supernova remnant Cassiopeia A
taken in three wavelength ranges (0.6-1.65 keV, 1.67-2.25 keV, and 2.25-7.5 keV) shown in
red, green, and blue and then combined into a single image.
+ + −→
Surface brightness:
We measure the luminosity ([erg s−1 ]) per area. This is often called Σ or I. It effectively
measures the magnitude per square arcsecond:
8
Images of supernova remnant created by Mark Vogelsberger using SAOImage DS9 image display and visualization tool for
astronomical data. https://sites.google.com/cfa.harvard.edu/saoimageds9/home
1. KEY OBSERVATIONS OF GALAXIES
Observationally, n ≈ 4 for ellipticals and n ≈ 1 for spirals. Theory needs to explain this!
– cluster environment
– spheroidal
– old stellar population
– pressure supported
– red color
M87
– de Vaucouleurs surface brightness profile
• Lenticular (SO):
– stellar disk
– no gas disk
– link between spiral and elliptical galaxies
NGC 2787
Galaxy luminosity distribution:
The luminosity L of an object is
Z
dE
L= = Iν dAdΩdν . (20)
dt
M87 Image: Courtesy of Canada-France-Hawaii Telescope / 9 NGC 2787 Image: NASA and The Hubble Heritage Team (STScI/AURA);
Coelum. Used with permission. Acknowledgment: M. Carollo (Swiss Federal Institute of Technology,
Zurich)
1. KEY OBSERVATIONS OF GALAXIES
What is the distribution function of L for galaxies? We commonly use the Schechter function
to describe the number density of galaxies at a given luminosity:
α
L dL
φ(L)dL = φ∗ e−L/L∗ (21)
L∗ L∗
mvc2 GM m
= (22)
r r2
for a circular orbit. This implies vc ∝ r−1/2 , for centralized mass, which is not constant. To
have vc constant, we need
1 r
Z
vc ∝ 4πr2 ρ(r)dr
r 0 (23)
−2
⇒ρ(r) ∝ r
to large radii. This was one of the first hints for dark matter.
What is dark matter? A few things we know:
• It can’t be non-luminous gas since we would have seen it through absorption lines
• Dim stars or other dense objects at larger distances (MACHOS: Massive Compact Halo
Object) have been ruled out since microlensing (the temporary brightening of a distant
object due to a closer massive object bending the light rays closer together) does not
occur frequently enough
• Neutrinos have been ruled out since they lead to the wrong structure formation because
they move so fast (hot dark matter). Since neutrinos move close to the speed of light,
they have too much kinetic energy to be bound in low-mass potential wells.
10
1. KEY OBSERVATIONS OF GALAXIES
• General theories:
– Cold Dark Matter (CDM): dark matter is a particle that moves slowly (v c)
and is collisionless, interacting solely through gravity.
– Self-interacting dark matter (SIDM): dark matter interacts through gravity as well
as through self-interactions that allow particles to scatter and transfer energy and
momentum.
– Warm dark matter (WDM): dark matter is still collisionless but moves with a
faster velocity than CDM, which makes it harder to form less massive halos.
– Bose-Einstein condensate (very low mass) dark matter: dark matter particles are
very low mass such that their de Broglie wavelength is on the length scale of
galaxies and leads to interference patterns in halos.
– Modified Newtonian dynamics (MOND): Dark matter is not a type of matter but
is accounted for through modifications to our theory of gravity.
Elliptical galaxies have a random motion velocity structure with velocity dispersion σv ∼
200 − 300 km/s. There is negligible circular motion, typically vc ∼ 50 − 100 km/s.
λobs − λ0 λobs
z= or 1+z = (24)
λ0 λ0
For low z (z 1), one finds that the distance d is related to the redshift through the
present-day Hubble constant H0 :
cz
d≈ , (25)
H0
which yields
v = H0 d ≈ cz , z 1 (26)
cz
The redshift directly yields the recession velocity. A more formal proof of d ≈ H0
will be
discussed later (low z limit for all distances).
11
2. STRUCTURE AND A QUALITATIVE PICTURE OF GALAXIES
spectrum of the galaxy. Galaxies are a combination of these, so the total flux at a given
frequency is a combination of the flux from each star:
Fν = NO Fν,O + NB Fν,B + ... (27)
Determining NO , NB , NA , NF ... is the basic idea of stellar population synthesis.
A galaxy is a collisionless fluid of stars and dark matter orbiting together with collisional gas
in a common self-gravitational potential.
With this definition, we can try to understand the main dynamical properties.
1 dI
G= . (32)
2 dt
Now consider the time derivative of G:
dG X ˙ X
= p~i · ~ri + p~ · ~r˙i
dt i i
X X
= F~i · r~i + mi vi2 (33)
i i
X
= F~i · ~ri + 2T
i
12
2. STRUCTURE AND A QUALITATIVE PICTURE OF GALAXIES
where T is the kinetic energy. Because gravity is a pairwise force, we can write
N
X
F~k = F~jk . (34)
i=1
Fii = 0 and 1 ≤ j ≤ N , so we can split Fjk into two parts, the upper and lower portions of
the matrix
k↓ j→
0 2
.. (35)
Fjk =
.
1 0
with
N X
X k−1 N
X −1 N
X
1 = F~jk · ~rk and 2 = F~jk · ~rk (36)
k=2 j=1 k=1 j=k+1
so
N
X X k−1
N X N
X −1 N
X
F~k · ~rk = F~jk · ~rk + F~jk · ~rk . (37)
k=1 k=2 j=1 k=1 j=k+1
which has the same matrix elements as the first term, so we get
N
X N X
X k−1
F~k · ~rk = F~jk · (~rk − ~rj ) . (40)
k=1 k=2 j=1
13
2. STRUCTURE AND A QUALITATIVE PICTURE OF GALAXIES
so we get
N
X N X
X k−1
F~k · ~rk = F~jk · (~rk − ~rj )
k=1 k=2 j=1
N Xk−1
X dV |~rk − ~rj |2
=− (42)
k=2 j=1
dr rjk
N X
k−1
X dV
=− rjk .
k=2 j=1
dr
n
We now assume the special case V (rjk ) = αrjk . This gives us
dV n−1
= nαrjk
dr (43)
dV
⇒ rjk = nV
dr
so
N
X N X
X k−1
F~k · ~rk = − nV (rjk )
k=1 k=2 j=1
X k−1
N X (44)
= −n V (rjk )
k−2 j=1
= −nVtot .
Finally:
dG X ~
= Ii · ~ri + 2T = 2T − nVtot (45)
dt i
dG 1 d2 I
With dt
= 2 dt2
,U = Vtot , and n = −1 (for gravity):
1 d2 I
= 2T + U . (46)
2 dt2
We now take the time average:
1 T dG
G(T ) − G(0)
Z
dG
= dt =
dt T T 0 dt T
(47)
dG
⇒ = 2hT iT − nhVtot iT
dt T
G(T )−G(0)
For a steady state system and long time average, T
≈ 0, so we get
0 = 2hT iT − nhVtot iT
. (48)
0 = 2hT iT + hU iT for n = −1
Note the three important assumptions for the virial theorem to hold:
14
2. STRUCTURE AND A QUALITATIVE PICTURE OF GALAXIES
• F is a pairwise force
2T + U = 0 . (49)
Assuming that a galaxy is made of N stars all with the same mass m (so total mass M = N m)
and average velocity ~v , we get a total kinetic energy for the system
1X 1 1
T = mi vi2 ≈ M v̄ 2 = M v 2 . (50)
2 i 2 2
From dimensional analysis for a galaxy of size R, we get a total potential energy
GM 2
U =− . (51)
R
The virial theorem then implies
GM 2
2
Mv + − =0
R
r (52)
GM
⇒v = .
R
Using some typical numbers:
R ≈ 10kpc M = 2 × 1033 g
2
11
M ≈ 10 M G = 0.0043M −1 pc kms
s
0.0043M −1 pc(km/s)2 1011 M
v=
10 000pc (53)
√
≈ 4 × 104 km/s ≈ 200 km/s
We can also use the virial to get the ideal gas law.
For an ideal gas with N particles at temperature T is
3
K = N kT (54)
2
where we use K for kinetic energy to differentiate from temperature and k is the Boltzmann
constant. The force comes from the pressure from the particles, so the force per unit area is
dF~ = −P dA
~. (55)
15
2. STRUCTURE AND A QUALITATIVE PICTURE OF GALAXIES
!
2.B Relaxation times
The virial theorem gave us some first insight into the dynamics of galaxies. Now we will
show that stars are collisionless, i.e. that two-body collisions are rare in galaxies. Since this
is true, we can describe the distribution of stars as a smooth density field and gravitational
potential.
Subject
star m x v
F⊥
r b
"
Field
star m
We assume that |δ~v |/|~v | 1 and that the field star is stationary. This means that δ~v is
perpendicular to ~v since the accelerations parallel to ~v cancel out as the subject star passes
by the field star. We calculate δv = |δ~v | by integrating F⊥ :
" 2 #−3/2
Gm2 Gm2 b Gm2 vt
F⊥ = 2 2
cos θ = 2 2 3/2
= 2 1+ . (60)
b +x (b + x ) b b
16
2. STRUCTURE AND A QUALITATIVE PICTURE OF GALAXIES
1 +∞
Z
δv = dtF⊥
m −∞
Gm +∞
Z
dt
= 2 2 i3/2
b
h
−∞
1 + vtb (61)
Gm +∞
Z
ds
=
bv −∞ (1 + s2 )3/2
2Gm
=
bv
using s = vtb . Thus, δv is roughly equal to the acceleration at closest approach, GM
b2
, times
the duration of the acceleration, 2b
v
.
Strong encounters:
An encounter is strong if δv ∼ v (which also causes the calculation to break down). This is
also when a star will have its path deflected by ∼ 90◦ ≡ bstrong .
GM
δv ∼ v ⇔ b . b90 = ≡ bstrong . (62)
v2
The cross section for strong encounters is
p = nσstrong R
N 4πR2
= 4 3 R (66)
3
πR N 2
3
= ∼ 10−11 .
N
This is a tiny probability! So there are likely no strong encounters in a galaxy. For globular
clusters, N ∼ 104 , so strong encounters are more common.
17
Subject
star m x v
2. STRUCTURE AND A QUALITATIVE PICTURE OF GALAXIES
F⊥
What about weakr encounters?
b
"
We have seen that strong encounters are rare, i.e. they practically never happen. Neverthe-
Field times, it will encounter many weak encounters. Each of
less, if a star crosses a galaxy many
star m until v ≈ v. The time it takes for this to happen is
those will slightly perturb its velocity ⊥
the relaxation time of the system.
log(ɸ)
Subject
! star m x v
F⊥
r b
"
Field
A star makes a random walk through a galaxy. Its total deviation from
star m
its path is the sum
of each of its encounters with other stars. For N encounters,
L* log(L)
N
X
2
δvtot = (δvi )2 . (67)
i=1
The strength of each encounter depends on the impact parameter b. The number of encoun-
ters N within (b, b + db) is
b+db
b
vΔt
We then have
X Z bmax
2
(δvi ) = (number of encounters in (b, b + db)) × (δv for each encounter with b)
i bmin
Z bmax 2
2Gm
= (2π v ∆t n b db)
bmin bv
bmax
8πG2 m2 n
Z
db
= ∆t .
v bmin b
(69)
18
2. STRUCTURE AND A QUALITATIVE PICTURE OF GALAXIES
bmax ≈ R ≈ 10 kpc
2R (70)
bmin ≈ bstrong = = 10−10 kpc (0.01AU)
v∗
so Z bmax
db bmax
= ln ≡ ln Λ (Coulomb logarithm)
bmin b bmin
10 kpc (71)
= ln
10−10 kpc
= ln 1011 ≈ 25 .
Relaxation time:
We define the relaxation time trelax through
X 8πG2 m2 n
(δvi )2 ≈ v 2 ⇒ ln Λtrelax
i
v
(72)
v3
⇒trelax = .
8πG2 m2 n ln Λ
We now compare this to the dynamical time torbit ≈ R/v of the system:
trelax v v4
=trelax = ,. (73)
torbit R 8πG2 m2 n ln ΛR
GM M/m
Using the virial theorem v 2 = R
and number density n = 4π 3
R
gives us:
3
(GM/R)2
=
8πG2 m2 M/m
4π 3 R ln Λ
R3
M
= 3
8πm 4π ln Λ
(74)
N N N
= = ∗ = ∗
6 ln Λ 6 ln bbmax R
6 ln 2R/N
min ∗
N∗
∼
6 ln N∗
which is very large! Thus, stars are orbiting in an unperturbed collective potential (colli-
sionless)!
19
particle loses
2. STRUCTURE AND A QUALITATIVE PICTURE OF GALAXIES
energy
and dark matter) must relax through a different process otherwise galaxies and galaxy clus-
ters would not reach a relaxed state within the age of the Universe. We say a system is
relaxed when its coarse grained phase-space distribution function does not change any more.
particle gains
Collisionless relaxation processes: energy
Phase mixing:
The coarse grained phase-space distribution function is distributed over time so doesn’t
change with time.
v v
x x
Changing
distribution function Constant
t=0 distribution function
t>0
Violent relaxation:
Since energy in the stellar and dark matter systems in galaxies can’t be efficiently exchanged
through collisions, we must find another way for energy exchange. The energy of an individ-
ual star (specific energy) is:
1
E = v2 + φ . (75)
2
Then the change in energy over time is
dE ∂E d~v ∂E dφ
= +
dt ∂~v dt ∂φ dt
~ + dφ
= −~v · ∇φ
dt
~ + ∂φ ∂φ d~x (76)
= −~v · ∇φ +
∂t ∂~x dt
~ + ∂φ ~
= −~v · ∇φ + ~v ∇φ
∂t
∂φ
=
∂t
Thus, the only way for a star to change its energy is by having a time-dependent potential.
To think of this intuitively, we can consider an object moving through a potential well. If
the potential is constant with time, the particle will recover the same energy as it comes out
20
Field
b+db
b log(L)
3. MODELLING GALAXIES
x x
the other side and there is no relaxation. If the potential grows with time, the particle will
Changing
need to expend more energy toDFcross it and will not have enoughConstant
energy to get back out of
DF
the potential well, thus losing energy. If the potential shrinks with time, the particle will
gain energy as it crosses the well.
no relaxation
particle loses
energy
particle gains
energy
As a galaxy or cluster forms, the gravitational potential changes significantly as mass accretes
and collapses into a halo. Averaging over all particles, the timescale for violent relaxation
tvr is
* 2 +−1/2
dE
dt
tvr =
E2
* +−1/2
∂φ 2
∂t (77)
=
E2
* +−1/2
φ̇2
∼
φ
where in the last step we used the time-dependent virial theorem (see Lynden-Bell 1967).
This occurs on roughly the same timescale as free-fall since this is the timescale at which
the potential changes during collapse. It’s very fast, hence ‘violent’ relaxation!
3 Modelling galaxies
So far, we have looked at the basic dynamical properties of galaxies. Now we discuss the
main ingredients of modelling galaxies:
21
3. MODELLING GALAXIES
Scalar potential:
~ = 1 F~
− ∇φ (78)
m
Note that mφ = U is the potential energy of the system and using Poisson’s equation
∇2 φ = 4πGρ, we get Z
ρ(~r) 3
φ(~r) = G d ~r (79)
|~r0 − ~r|
⇒ potential φ − density ρ − pairs! (80)
Examples:
• Kepler/point mass potential:
GM
φ=− (81)
r
To find F~ , we take the gradient of φ
1 ~ GM
F = 2 êr . (82)
m r
• Homogeneous sphere:
M
ρ(~r) = 4
3
πR3 (83)
F~ =?
We do not have φ, so we need a different way to get F~ . We can use Gauss’s theorem
for gravity for a surface Sr with radius R enclosing a volume Vr :
Z Z
~ ~
F · dS = ~ · F~ )dV
(∇
Sr Vr
Z
= −m (∇ ~ 2 φ)dV
V Z (84)
= −4πGm ρ(~r)dV
22
3. MODELLING GALAXIES
" r
r 1 + | cos θ|
φ(r, θ) = vc2 ln + ln (89)
r0 2
Is this a disk? It’s hard to see based on the potential, so we need to find ρ. Let’s look
at Poisson’s equation:
∂ 2φ
2 1 ∂ 2 ∂φ 1 ∂ ∂φ 1
∇ φ= 2 r + 2 sin θ + 2 2
r ∂r ∂r r sin θ ∂θ ∂θ r sin θ ∂ϕ (90)
| {z }
= 0, since no ϕ dependence
Using φ = vc2 φ0 :
vc2 ∂ vc2 ∂ 2 φ0
2 21 ∂φ0
∇ φ= 2 r + 2 cos θ + sin θ 2
r ∂ r r sin θ ∂θ ∂θ
2
2
(91)
v cos θ ∂φ0 ∂ φ0
= c2 1 + +
r sin θ ∂θ ∂θ2
2
We now calculate ∂φ∂θ
0
and ∂∂θφ20 . We assume cos θ > 0. The calculations are the same
or cos θ < 0 except for an overall sign change cos θ → − cos θ.
r 1 + cos θ
φ0 = ln + ln (92)
r0 2
23
3. MODELLING GALAXIES
Then
∂φ 2 sin θ sin θ
= − = −
∂θ 1 + cos θ 2 1 + cos θ
2 2 (93)
∂ φ0 cos θ sin θ
= − −
∂θ2 1 + cos θ (1 + cos θ)2
So
cos θ ∂φ0 ∂ 2 φ0 2 cos θ sin2 θ
+ = − −
sin θ ∂θ ∂θ2 1 + cos θ (1 + cos θ)2
−2 cos θ − 2 cos2 θ − sin2 θ
=
(1 + cos θ)2
1 + cos2 θ + 2 cos θ (94)
=−
(1 + cos θ)2
(1 + cos θ)2
=−
(1 + cos θ)2
= −1
2
For cos θ 6= 0, this gives ∇2 φ = vrc2 (1 − 1) = 0, so there is
no density for θ 6= π/2 and all mass is in a thin plane with
infinite density ρ (3D density).
24
3. MODELLING GALAXIES
When θ > π2 , cos θ < 0 and | cos θ| = − cos θ, and when θ < π2 , cos θ > 0 and
| cos θ| = cos θ. So we take the derivative using − cos θ in the first term and cos θ in
the second term
!
1 sin θ − sin θ
Σ(r) = −
4πG 1 − cos θ π + 1 + cos θ π −
2 2
( → 0)
(98)
1 vc2
= (1 + 1)
4πG r
1 vc2
⇒ Σ(r) = .
2πG r
Simulations showed the ρ0 and a are strongly correlated for CDM halos, so halos are
approximately members of a 1-parameter family. The conventional choice for this
25
3. MODELLING GALAXIES
parameter is r200 , the distance which has an enclosed density 200 times the cosmic
critical density ρc (which we will cover later) or M200 = 200ρc 43 πr200
3
.
Related topics:
– Core-cusp problem: From observations of stellar dynamics, the inner profile of
halos flattens to a slope ∼ 0 (core) instead of −1 (cusp). This is possibly due to
supernova feedback, but it could also be resolved through modifications of cold
dark matter.
– Diversity of shapes problem: Observationally, halos display diversity in the shapes
of their profiles with some cuspier and some more cored profiles whereas, in simu-
lations, halos are universally described by the NFW profile and self-similar across
mass ranges (the profiles look the same when scaled).
– Missing satellite problem: Simulations produce more satellite halos than there are
observed satellite galaxies. It’s possible that not all subhalos form stars, so we
need to be able to find “dark subhalos." This could be done by looking for disrup-
tions in stellar streams or through gravitational lensing. Recently, however, there
have been many more satellites found as our observational techniques improve.
– Too-big-to-fail problem: This is related to the missing satellites problem, where
the number of predicted large halos doesn’t match the number of large galaxies
observed (but the total number of satellite halos is consistent). The gravitational
potential of these galaxies, however, is large enough that they should have col-
lected enough gas and stars to form galaxies and maintain their evolution (e.g.
not lose the stars through stripping).
3.B Orbits
Now that we have looked at potential-density pairs, we can study orbits in these potentials.
Orbits refer to the motion of stars through 6D phase space (~x(t), ~v (t)). Often, the integrals
of motion restrict the dimensionality of the orbit (1 per integral of motion).
Integrals of motion:
The orbital energy E is:
1 1
E = v 2 + φ(r) = ṙ2 + φ(r) (102)
2 2
26
3. MODELLING GALAXIES
dE dφ
= ṙr̈ + ṙ = ṙr̈ − r̈ṙ = 0 (103)
dt dr
which implies that the energy is constant along the orbit.
~ (for a central force potential) is:
The angular momentum L
~ = ~r × ~r˙
L (104)
cos ψ
êr =
sin ψ
(106)
sin ψ
êψ =
− cos ψ
We have:
d d
~r = (rêr )
dt dt
d
= ṙêr + r (êr )
dt !
dêr dr dêr dψ
= ṙêr + r +
dr dt
|{z} dψ dt
=0 (107)
d
= ṙêr + rψ̇ êr
dψ
| {z }
d cos ψ − sin ψ
= = = êψ
dψ sinψ cosψ
= ṙêr + rψ̇êψ
27
3. MODELLING GALAXIES
so
d2 d
2
~r = (ṙêr + rψ̇êψ )
dt dt
d
= (r̈êr + ṙψ̇êψ ) + (rψ̇êψ )
dt
d
= r̈êr + ṙψ̇êψ + ṙ(ψ̇êψ ) + r ψ̇êψ
dt
d
= r̈êr + ṙψ̇êψ + ṙψ̇êψ + rψ̈êψ + rψ̇ êψ
| {zdt } (108)
dêψ dψ
=
dψ dt
= −ψ̇êr
= r̈êr + ṙψ̇êψ + ṙψ̇êψ + rψ̈êψ − rψ̇ 2 êr
= r̈ − rψ̇ 2 êr + 2ṙψ̇ + rψ̈ êψ
and, since we are using a central force,
d2
~r = F (r)êr (109)
dt2
where F is the force per unit mass. Combining the two above equations, we get the scalar
equations for 4D orbits for the radial and tangential components of motion:
radial : r̈ − rψ̇ 2 = F (r)
(110)
tangential : 2ṙψ̇ + rψ̈ = 0 .
1
For now, we focus on the radial equation and substitute u = r
to avoid the singularity at
r = 0. Then 2
d2 1
1 dψ
F (r) = 2 − . (111)
dt u u dt
With
~ = ~r × ~v ⇒ L = r2 dψ
L (112)
dt
we can parameterize t with ψ to get u = u(ψ):
d L d d
= 2 = Lu2 (113)
dt r dψ dψ
Then
~ = ~r × d ~r
L
dt
= ~r × ṙêr + rψ̇êψ
= rêr × ṙêr + rψ̇êψ (114)
= r2 ψ̇
dψ
= r2 .
dt
28
3. MODELLING GALAXIES
Then we get
d
2 2 d 1 1 2 dψ
F (u) = Lu Lu − Lu
dψ dψ u u dψ
d −1 du 1 2
= Lu2 Lu2 2 − Lu2 (115)
dψ u dψ u
2
du
= −L2 u2 2 − L2 u3
dψ
which gives us the orbit equation:
d2 u F (u)
2
+u=− 2 2 (116)
dψ Lu
Examples:
• Kepler:
GM d GM
φ(r) = − → F (r) = − φ = − 2 = −GM u2 (117)
r dr r
so we get the orbital equation:
d2 u F (u) GM
2
+u=− 2 2 = 2 . (118)
dψ Lu L
Note that this is a harmonic oscillator, so we know the solution:
GM
u(ψ) = C cos(ψ − ψ0 ) + (119)
L2
Then for ψ = 0 to ψ = 2π, u(ψ = 0) = u(ψ = 2π) and we get closed orbits (frequency
ω = 1).
29
3. MODELLING GALAXIES
d2 u F (u) GM 4G2 M 2
+ u = − = + u
dψ 2 L2 u2 L2 L2 c2
!
d2 u 4G2 M 2 GM (122)
⇒ 2 + 1− 2 2
u= 2 .
dψ | L{zc } L
constant
The term in parentheses implies that k 2 6= 1, so ω < 1, which means that u(ψ = 0) 6=
u(ψ = 2π). This accounts for the precession of Mercury.
The solution for a harmonic oscillator
is
u = cos(ωt − φ) (124)
where ω 2 = k/m. Then if w 6= 1 and k 6= 1, u(0) 6= u(2π). Here, t is analogous to ψ,
so if the position u after one orbit when ψ = 2π is not the same as when ψ = 0, the
mass has not returned to its previous position and the orbit is not closed.
30
3. MODELLING GALAXIES
Note that
d
êr = ψ̇êψ
dt
d
êψ = −ψ̇êr (126)
dt
d
êz = 0
dt
then
d2 d
= ṙê r + r ψ̇êψ + żêz
dt2 dt
d
= r̈êr + ṙψ̇êψ + ṙψ̇êψ + r ψ̇êψ + z̈êz
dt
= r̈êr + ṙψ̇êψ + ṙψ̇êψ + rψ̈êψ − rψ̇ 2 êr + z̈êz (127)
= r̈ − rψ̇ 2 êr + 2ṙψ̇ + rψ̈ êψ + z̈êz
| {z }
1d
2
= r ψ̇
r dt
and for an axisymmetric potential
d2
~ ∂φ ∂φ
~r = F = − , 0, − (128)
dt2 ∂r ∂z
so we get each component of the force:
∂φ
radial : r̈ − rψ̇ 2 = −
∂r
1d 2
tangential : r ψ̇ = 0
r dt
d 2 (129)
⇒ r ψ̇ = 0
dt
(using conservation of Lz = r2 ψ = constant)
∂φ
vertical : z̈ = − .
∂z
We then rewrite this in terms of the effective potential:
L2z
φeff = φ + (130)
2r2
where the last term is the centrifugal barrier. Since
d
~v = ~r (131)
dt
then, using the above from ~r˙ ,
1 2
E= ṙ + r2 ψ̇ 2 + ż 2 + φ
2 (132)
1 2
ṙ + ż 2 + φeff
=
2
31
3. MODELLING GALAXIES
so
∂φ
r̈ = rψ̇ 2 =
∂r
L2z
2 ∂
= rψ̇ − φeff − 2
∂r 2r
2
∂ψeff L (133)
= rψ̇ 2 − − 3z
∂r r
4 2
∂φ eff r ψ̇
= rψ̇ 2 − − 3 (using Lz = r2 ψ̇)
∂r r
∂φeff
=−
∂r
and
∂φeff
z̈ = − . (134)
∂z
So finally we get the scalar equations for 4D orbits (E, Lz ):
∂φeff
r̈ = −
∂r
∂φeff
z̈ = − (135)
∂z
1 2
ṙ + ż 2 + φeff .
E=
2
Note that the orbits have a uniform rotation around the symmetry# r #=-π/2
axis (z) with ψ̇ = Lr2z , #=π/2
but we have oscillations in r and z. If r is not oscillating, then z = 0, and any perturbation ê$
leads to oscillations in z and r. êr
r
$
Guiding center and circular orbits:
φeff has a minimum at some Rg such that for a given Lz , φ(Rg , 0) is minimal:
At the minimum:
∂φeff ɸeff(r,0)
=0
∂r r=Rg ,z=0
(136)
∂φeff
=0
∂z r=Rg ,z=0 Rg
We also have
L2z
Ω2 = ψ̇ 2 = . (139)
Rg4
And since
∂φeff ∂φeff
= =0 (140)
∂r Rg ,0 ∂z Rg ,0
then
r̈ = z̈ = 0 (141)
so we have a circular orbit with speed Ω = ψ̇. The minimum of φeff occurs at a radius Rg
at which a circular orbit has angular momentum Lz and E = φeff . This orbit is called the
guiding center. If an object is pushed off the guiding center, there is a restoring force that
leads to oscillations, or epicycles.
Epicycle approximation:
In disk galaxies, many stars are on mostly circular orbits, but they are not exactly circular.
We look for small perturbations around the circular orbit.
We will define our coordinate system (x, y) as
x = r − Rg
(142)
y=z
∂ φ̃eff
φ̃eff,y = =0 (144)
∂y 0,0
∂ 2 φ̃eff
φ̃eff,xy = =0 (by symmetry).
∂x∂y 0,0
We define κ and ν:
∂ 2 φeff
κ2 ≡ φ̃eff,xx = φeff,rr =
∂r2 Rg ,0
2
(145)
2 ∂ φeff
ν ≡ φ̃eff,yy = φeff,zz = .
∂z 2 Rg ,0
so
1 1
φ̃eff ≈ κ2 x2 + ν 2 y 2 + φ̃eff (0, 0)
2 2 (146)
1 1
= (κx) + (νy)2 + φ̃eff (0, 0) .
2
2 2
33
3. MODELLING GALAXIES
ẍ = −κ2 x
. (149)
ÿ = −ν 2 y
This is harmonic oscillation with epicycle frequency κ and vertical frequency ν in addition
to the circular frequency r
Lz vc 1 ∂φ
Ω(r) = 2 = = (150)
r r r ∂r
where for the last equality we used the fact that the centripetal force is equal to the gravi-
2
tational force vrc = ∂φ
∂r
on a circular orbit.
34
3. MODELLING GALAXIES
2 L2z
=Ω = 4
r
L2z ∂ 2 φ L2z
= − 4 + 2 +4 4
r ∂r r Rg
2
∂ φ L 2
(151)
= 2
+ 3 4z
∂r r Rg
2
∂ ∂φ Lz
= − 3
∂r ∂r r Rg
2
∂ ∂ L
= φ + z2
∂r ∂r 2r Rg
| {z }
= φeff
2
∂ φeff
=
∂r2 Rg
=κ2
d(Ω2 )
2 2
⇒κ = r + 4Ω .
dr
35
r
ê$
# r #=-π/2 #=π/2 êr
3. MODELLING GALAXIES r
ê$ $
êr
We then integrate over time, so r
2Ωg r0
$
ψ(t) = Ωg t + ψ0 − sin(κt + α) . (154)
κRg
ɸeff(r,0)
We use the galactic coordinate system to measure the location of stars in the sky (l, b):
b l: galactic longitude
l b: galactic latitude
sun R galactic
center
R is the distance from the sun to the galactic center (∼ 8kpc) and d is the distance from the
sun to the star. l measures the angle in the plane of the Milky Way away from the line of
sight to the galactic center, and b measures the angle above the plane of the galaxy. Within
this system, we find:
proper motion : µ ≈ d(A cos(2l) + B)
(155)
line of sight motion : vk ≈ dA sin(2l)
where A and B are the Oort constants given by:
1 dΩ
A=−
2dR (156)
1 dΩ
B =− Ω+ R .
2 dR
More importantly, they can be related to κ:
κ2 = −4B(A − B) . (157)
36
3. MODELLING GALAXIES
Luminosity-velocity relations:
We can relate properties of a galaxy to observables through several equations:
R
θ= (apparent size)
d
L
F = (158)
4πd2
GM
v2 = .
R
Introducing surface brightness Σ
F L d2
Σ= = ·
θ2 4πd2 R2 (159)
L v4
= · 2 2
4π G M
then
v4
L= . (160)
Σ4πG2 (M/L)2
If we assume, for a given class of galaxies, that the surface brightness and the mass-to-light
ratio are the same, then
L ∝ v4 . (161)
This introduces two important relations.
The Tully-Fischer relation is used for spiral galaxies and relates the maximum velocity in
the rotation curve vmax , which can be measured from HII spectra, and the luminosity:
4
L ∝ vmax . (162)
The Faber-Jackson relation is used for ellipticals and relates the velocity dispersion σv to the
luminosity:
L ∝ σv4 . (163)
Thus, we can get an estimate of the intrinsic luminosity of a galaxy be measuring stel-
lar velocities. The constant of proportionality is roughly L∗ /(220 km/s)4 , where L∗ is the
characteristic galaxy luminosity.
37
3. MODELLING GALAXIES
w
~ = (~x, ~v ) (166)
then
∂f ∂ ˙
+ fw~ =0. (167)
∂t ∂w~
This is the same form as the standard 3D continuity equation. We can rewrite this by
~ and using velocity ~v = ~x˙ and acceleration ~a = ~v˙ :
expanding out w
∂f ∂ ˙
0= + fw ~
∂t ∂w~
∂f ∂ ∂
= + (f ~x˙ ) + (f ~v˙ )
∂t ∂~x ∂~v (168)
∂f ∂ ∂ ~
= + (f~v ) + f (−∇φ)
∂t ∂~x ∂~v
∂f ∂f ∂φ ∂f
= + ~v − .
∂t ∂~x ∂~x ∂~v
This gives us the collisionless Boltzmann equation (CBE):
∂f ∂f ∂φ ∂f
+ ~v − =0 . (169)
∂t ∂~x ∂~x ∂~v
df
Note that another way to see this is by writing out dt
= 0 and taking the limits lim~x→∞ = 0
and lim~v→∞ = 0.
Moments of the phase-space density give us some average quantities of the system.
38
3. MODELLING GALAXIES
We now examine moments of the collisionless Boltzmann equation more closely. We break
each integral into three terms to simplify each individually.
a) First moment: Z
∂f 3 ∂f ∂φ ∂f
d ~v + ~v − =0
∂t ∂~x ∂~x ∂~v
Z Z Z
3 ∂f 3 ∂f ∂φ ∂f (173)
d ~v + d ~v ~v − d3~v =0
∂t ∂~x ∂~x ∂~v
| {z } | {z } | {z }
1 2 3
Z Z
∂f3 ∂ ∂n
1 : d ~v = d3~v f =
∂t ∂t ∂t
Z Z
∂f ∂ ∂ X ∂
2 : 3
d ~v ~v = 3
d ~v ~v f = n~v¯ = (nv̄i ) (174)
∂~x ∂~x ∂~x i
∂x i
Z Z
3 ∂φ ∂f ∂φ ∂f ∂φ ~v=+∞
3 : d ~v = d3~v = [f ] =0
∂~x ∂~v ∂~x ∂~v ∂~x ~v=−∞
For the third term, we used the fact that phase-space distribution goes to 0 at ±∞ for
physical systems.
∂n ∂
n~v¯ = 0 .
+ (175)
∂t ∂~x
b) Second moment:
Z
3 ∂f ∂f ∂φ ∂f
d ~v vj + ~v − =0
∂t ∂~x ∂~x ∂~v
Z Z Z
∂f ∂f ∂φ ∂f (176)
d3~v vj + d3~v vj ~v − d3~v vj =0
∂t ∂~x ∂~x ∂~v
| {z } | {z } | {z }
1 2 3
39
3. MODELLING GALAXIES
Z Z
3 ∂f ∂ ∂ ∂n ∂v̄j
1 : d ~v vj = d3~v vj f =
(nv̄j ) = v̄j + n
∂t ∂t ∂t ∂t ∂t
X ∂ ∂v̄j ∂v̄j X ∂
= −v̄j (nv̄i ) + n =n − v̄j (nv̄i )
i
∂x i ∂t ∂t i
∂x i
∂n X ∂
using the continuity equation =− (nv̄i )
∂t i
∂x i
Z Z Z
3 ∂f 3
X ∂f X ∂
2 : d ~v vj ~v = d ~v vj vi = d3~v vj vi f
∂~x i
∂xi i
∂xi
| {z }
2
= nvj vi = n σij + v̄i v̄j
X ∂
n σij2 + v̄i v̄j
=
i
∂x i
Z Z X ∂φ Z
3 ∂φ ∂f 3
X ∂φ ∂f ∂f
3 : d ~v vj = d ~v vj = d3~v vj (177)
∂~x ∂~v i
∂xi ∂vi i
∂xi ∂vi
((k, l, i) are permutations of (1, 2, 3))
X ∂φ Z Z Z
∂f
= dvk dvl dvi vj
i
∂xi ∂vi
| {z }
Z
vi =+∞ ∂vj
= [vj f ]vi =−∞ − dvi f
∂vi
Z
= 0 − dvi δij f
X ∂φ Z Z Z
=− dvk dvl dvi δij f
i
∂xi
X ∂φ Z
=− d3~v δij f
i
∂xi
∂φ
= −n
∂xj
40
3. MODELLING GALAXIES
To simplify, we take the radial Jeans equation and focus on steady-state symmetric systems.
Implications:
∂
• ∂t
= 0 since we have steady state
41
3. MODELLING GALAXIES
σt2
β =1− 2
(184)
σrr
This depends only on radial components with uncertainty from β, assuming spherical sym-
metry and a steady-state system.
r2 1 ∂ 2
2 2βσrr
M (< r) = − (nσrr ) +
G n ∂r r
2
rσrr r ∂ 2
=− 2 ∂r
(nσrr ) + 2β
G nσrr
2
2
(186)
rσrr r dn r dσrr
=− + 2 + 2β
G n dr σrr dr
2 2
rσrr d ln n d ln σrr
=− + + 2β
G d ln r d ln r
42
d
b
l
3. MODELLING GALAXIES sun R galactic
center
overdensity
Consider a nearly uniform distribution of stars with per-
turbations with respect to a static uniform background.
Guiding
We can study the stability of this configuration by in- Center
specting the continuity and the Jeans equations.
non-rotating
static system
We can rewrite the number density n using ρ = mn and assume that σij is isotropic so the
pressure is P = ρσij2 = ρσij = mnσij2 . Then we can rewrite the Jeans equations as:
∂~v ~ ~ − 1 ∇P
~ .
+ ~v · ∇ ~v = −∇φ (188)
∂t ρ
Similarly, the continuity equation becomes:
∂ρ ~
+ ∇ · (ρ~v ) = 0 . (189)
∂t
Note that we have dropped the ¯ ’s (average value symbols) for simplicity in our
equations and ~v is referring to the average velocities at (~x, t). We will continue with this
convention in the following calculations.
Small perturbations:
For a small perturbation in a static uniform background, we have
ρ = ρ0 + ρ1 (~x, t)
~v = ~v0 + ~v1 (~x, t)
(190)
P = P0 + P1 (~x, t)
φ = φ0 + φ1 (~x, t) .
We can choose φ0 = 0 and, since the background is static, ~v0 = ~0. ρ0 and P0 are both
nonzero constants. Note that this is not a physical set of conditions since Poisson’s equation
gives ∇2 φ0 = 4πGρ0 so φ0 = 0 implies ρ0 = 0, but we continue with our calculations ignoring
this. This is known as the Jeans swindle.
43
3. MODELLING GALAXIES
∂ 2 ρ1 ~ 2 ρ1 = 0 .
− 4πGρ0 ρ1 − vs2 ∇ (199)
∂t2
44
3. MODELLING GALAXIES
which gives the time evolution of perturbations. We plug this into the wave equation and
get
w2 = vs2 k 2 − 4πGρ0 . (201)
We have two solutions:
If w = 0: 2
πvs2
2π
λ2J = = . (202)
k Gρ0
Jeans length and mass:
The Jeans length λJ is the maximum size a perturbation can be to remain stable. The Jeans
mass MJ is the corresponding mass enclosed within the Jeans length of a given substance.
πvs2 πσ 2
λ2J
= =
Gρ0 Gρ0 . (203)
4
MJ = πρ0 λ3J
3
So we have stability for λ < λJ and M < MJ . Note that for collisional gas, the Jeans length
is determined by the sound speed vs and for collisionless dark matter and stars, the Jeans
length is determined by the pressure from the velocity dispersion σ.
45
r
Rg
3. MODELLING GALAXIES d
b
which is similar to the Jeans length result, differing only
l by a factor of π. So, random motion
and pressure can stabilize perturbations on small scales.R galactic
sun
center
During the perturbation, R → R0 with R0 = R − dr. We want to know when this will
lead to collapse and when it will be stable. This is a competition between centripetal and
gravitational forces.
46
3. MODELLING GALAXIES
Full stability criterion: On small scales, we have stability if R < λJ and on large scales,
we have stability if R > Rrot . Small scales are stabilized by random motion and large scales
are stabilized by rotational motion. The system is unstable if λJ < R < Rrot . We can
combine the two criteria and get full stability when λJ ≥ Rrot . This gives us (adapting λJ
from an arbitrary 3D potential to a 2D disk):
π σ2 2πGΣ
≥
8 GΣ 3Ω2 (212)
4 GΣ
⇒σcrit ≥ √ .
3 Ω
Note that the angular speed of the patch is only approximately Ω. It actually rotates with
epicyclic frequency κ, which is not too far off from Ω for real galaxies. We can relate κ to Ω
for a typical galactic disk:
dΩ2
2 2
κ (Rg ) = r + 4Ω Rg
(213)
dr
and in galaxies with circular velocity that is approximately constant:
vc √
Ω= ⇒ κ2 = 2Ω2 ⇒ κ = 2Ω . (214)
r
So for galaxies: r
4 GΣ 32 GΣ
σcrit ≥ √ √ = .
3 κ / 2 | {z3} κ
2
(215)
3.26
Toomre finds
GΣ
σcrit = 3.26 (216)
κ
Toomre criterion Q:
We can write the stability criterion Q for rotating disks:
σ > 1 : stable
Q= (217)
σcrit < 1 : unstable
Extended Data Figure 7b. Neeleman, M., Prochaska, J.X., Kanekar, N. et al. Code used to generate kinematic models: mneeleman, and J. Xavier
A cold, massive, rotating disk galaxy 1.5 billion years after the Big Bang. 47 Prochaska. “Mneeleman/qubefit: Small Documentation Updates”.
Nature 581, 269–272 (2020). https://doi.org/10.1038/s41586-020-2276-y Zenodo, February 11, 2021. https://doi.org/10.5281/zenodo.4534407.
3. MODELLING GALAXIES
• Stars are born with a certain mass spectrum. This is the initial stellar mass function
(IMF) φ(m)
The galactic spectrum is a superposition of stellar spectra. Adding the stellar spectra taking
into account ψ(t) and φ(m) allows us to constrain the initial mass function and star formation
rate of galaxies. We can use this to learn about the stellar population and galaxy evolution.
dm
(t) = . (218)
dt
There are a few observational indications for the star formation rate:
SFRFIR LFIR
∼ (219)
M /yr 5.8 × 109 L
SFRUV LUV
∼ (221)
M /yr 7.2 × 1027 erg/s
For a given galaxy, the star formation rate depends on the density and temperature of the
gas. When gas is cold and dense, it is able to collapse into stars. The star formation rate
can be roughly approximated by dividing the gas mass by the free-fall time.
48
3. MODELLING GALAXIES
Torrey, P., M. Vogelsberger, et al. MNRAS. 484, no. 4(2019): 5587-5607. © Oxford M. Krumholz et al 2009 ApJ 699 850. American Astronomical Society (AAS). All rights
M.R.
University Press. All rights reserved. This content is excluded from our Creative reserved. This content is excluded from our Creative Commons license. For more
Commons license. For more information, see https://ocw.mit.edu/help/faq-fair-use/ information, see https://ocw.mit.edu/help/faq-fair-use/
8 Gas Mass
Hot : 7. 5%
7
6
Log(T[K])
WHIM : 49. 5%
4
Diffuse : 38. 7% Cond : 4. 3%
38 6 4 2 0 2
Log(ρ[cm −3 ])
The above plot on the left is from Torrey et al. 2019 shows a phase diagram of gas in the
IllustrisTNG simulations at z = 0 and is split into various phases of the ISM. Darker regions
show a higher gas mass, and the percentages show the total fraction of gas mass in each
phase. The condensed material is in the lower right corner, and stars form along the thin
line. The plot to the right is from Krumholz et al. 2009 shows the star formation rate surface
density as a function of gas surface density. Each point is a different galaxy, compiled from
several sources (different colors).
The star formation rate of a galaxy depends primarily on the molecular gas within a galaxy
rather than the total gas. However, it can be difficult to predict what fraction of a galaxy’s
gas is in the molecular phase. This fraction depends on the total gas density, metallicity,
and clumping on small scales. To get more precise predictions for star formation, it is also
necessary to consider events such as supernovae and shocks.
is normalized so Z mh
mφ(m)dm = 1M (223)
ml
ml ∼ 0.08M since hydrogen fusion can’t occur in stars lower than this and mh ∼ 100M
since the Eddington limit prevents stars larger than this.
Example:
M∗ is the total mass of newly formed stars. Then the total number dN (m) and total mass
dM (m) of stars born in (m, m + dm) are
M∗
dN (m) = φ(m)dm
M
(224)
M∗
dM (m) = mφ(m)dm .
M
Stellar spectra:
The stellar spectrum of a star is given by its luminosity L, effective temperature Teff , and
chemical composition z. The evolution of a star in the (L, Teff ) plane (stellar evolutionary or
HR diagram track) only depends on the initial mass and initial metallicity. Once the initial
mass and metallicity are known, one can calculate a stellar spectrum.
Initial mass function plot by JohannesBuchner on Wikimedia Commons. License CC-BY-SA. © JohannesBuchner 2015. All rights reserved. This content is
excluded from our Creative Commons license. For more information, see https://ocw.mit.edu/help/faq-fair-use/
50
Guiding
Center
3. MODELLING GALAXIES
non-rotating rotating
static system non-static system
Population synthesis:
A galaxy spectrum is a superposition of stellar spectrum:
Z t
Lλ = Lcp 0 0 0 0
λ (t − t , Z(t ))ψ(t )dt . (230)
0
We can measure time from t0 that the stars formed so τ = t − t0 and τ0 = t0 . ψ(t0 ) is the star
formation rate at t0 . Lλ is the luminosity at λ per unit stellar mass of all stars of a coeval
population of age τ with initial metallicity Z(τ0 ):
Z
cp φ(m)
Lλ (τ, Z(τ0 )) = Lλ (m, Z(τ0 ), τ ) dm (231)
M
where Lλ is the luminosity at wavelength λ of a star with initial mass m and initial metallicity
Z(τ0 ) at time τ .
A few notes:
• Lλ (t) is a convolution of φ, ψ, and Lλ .
• φ and ψ are not known precisely.
• There are sophisticated codes available to numerically iterate to figure out φ and ψ:
– assume an initial mass function φ
– impose a star formation rate
– run convolution
– compare with data (can break some degeneracies with spectral features)
– adjust SFR and IMF and repeat.
51
3. MODELLING GALAXIES
Note that the metallicity measures by mass and abundance measures by number.
This forms a complete chemical model. We can figure out individual terms and then solve.
We will use some approximations for an analytical solution.
Z ∞
E(t) = (m − wm )ψ (t − τMS (m)) φ(m)dm (235)
mt
mt : Main sequence turnoff mass. This is the lowest mass of stars dying at time t.
t−τm (m) φ(m): Birth rate of stars of mass m at time t − τm (m), which is the death rate
at time t.
Z ∞
Ez (t) = [(m − wm )Z(t − τMS (m)) + mρZm ] ψ(t − τMS (m))φ(m)dm (236)
mt
(m − wm )z(t − τMS (m)): mass of metals that at time t − τMS (m) were locked in a star
of mass m and are now ejected with the envelop at time t.
mρZm : new metals produced by a star of mass m. (Note: some elements get destroyed,
for example lithium has a ρzm < 0.)
Stars below mlim never lose mass while stars greater than mlim immediately lose mass. This
is because massive stars get off the main sequence so quickly (τMS ∝ M −3 ) so this process is
essentially instantaneous, τMS ≈ 0. R is the returned mass per star formed:
Z ∞
R= (m − wm )φ(m)dm . (239)
mlim
Then Z ∞
Ez (t) ≈ [(m − wm )Z(t) + mρZm ] φ(m)dm
mlim
Z ∞
(240)
= ψ(t)Z(t)R + ψ(t) mρZm φ(m)dm
mlim
= ψ(t)Z(t)R + (1 − R)yψ(t)
53
3. MODELLING GALAXIES
where y is the mass of produced metals per remnant mass (white dwarfs, neutron stars, etc.)
Z ∞
1
y= mρZm φ(m)dm . (241)
1 − R mlim
This gives us the equations of chemical evolution in the instantaneous recycling approxima-
tion:
M = Ms + Mg
dM
=f −e
dt
dMs
= (1 − R)ψ(t)
dt (242)
dMg
= −(1 − R)ψ(t) + f − e
dt
d(ZMg )
= −zψ + RZ(t)ψ(t) + (1 − R)yψ(t) + Zf f − Ze
dt
= (1 − R)(−Z + y)ψ + Zf f − Ze
dMg d(ZMg )
We can combine the last two equations for dt
and dt
to get
dZ
Mg = (1 − R)yψ(t) + (Zf − Z)f + Ze . (243)
dt
Closed-box model:
The simplest evolution model is to assume a closed box (f = e = 0) containing only gas
(Mg (0) = M, Ms (0) = 0 with zero metallicity (Z(0) = 0)). The equations then simplify:
M = Ms + Mg
dM
=0
dt
dMs
= (1 − R)ψ(t)
dt
dMg (244)
= −(1 − R)ψ(t)
dt
d(ZMg )
= (1 − R)(−Z + y)ψ
dt
dZ
Mg = (1 − R)yψ(t) .
dt
dMg
We can divide dt
by Mg dZ
dt
to get
1 dMg 1
=− (245)
Mg dZ y
and integrate Z Z(t)
Mg (t) Mg (t) dZ Z(t)
ln(Mg ) M
= ln = − =−
M 0 y y
(246)
M
⇒Z(t) = y ln
Mg (t)
54
3. MODELLING GALAXIES
so we have:
Mg (t = 0)
Z(t) = y ln (247)
Mg (t)
which is the metallicity of gas as a function of only Mg (t).
Metallicity of stars:
In the closed box model, stars and gas must contain all the metals ever produced.
Z tZ ∞
Zs Ms + ZMg = mρZm ψ(t0 )φ(m)dmdt0 (248)
0 0
where Zs is the average metallicity of stars and the right side of the equation represents the
mass of all metals injected into the interstellar medium until time t. Then
Z t
Zs Ms + ZMg = (1 − R)yψ(t0 )dt0 ≈ (1 − R)y ψ̄(t)t . (249)
0
Mg u
Zs = y − Z =y−Z . (252)
M − Mg 1−u
So we finally get
1
gas : Z(s) = y ln
u(t)
(253)
u(t)
stars : Zs (t) = y − Z(t)
1 − u(t)
where u is the gas fraction Mg /M and M is constant for the closed box model. As u →
0, Zs → y, which gives the typical metallicity of stars. This must be less than or equal to
the typical yield.
G-dwarf problem:
We want to measure the metallicity distribution of G stars. These stars have not evolved
much and are still on the main sequence. Their age is so high that they formed from a
very low metallicity gas, since Z(0) = 0. We can use the closed box result to predict their
metallicity distribution. However, we cannot use an average Zs since we are looking for a
distribution.
55
3. MODELLING GALAXIES
Ms (≥ u) 1−u
Ms (≥ u) = (1 − u)M ⇒ = (254)
Ms,0 1 − u0
where Ms (≥ u) is the mass of stars formed while the gas fraction was ≥ u and ...0 refer to
present-day values.
Then the stellar mass fraction Ms (≥ u)/M was made from gas with Z ≤ y ln(1/u) and we
can rewrite the stellar mass fraction using u = e−Z/y and u0 = e−Z0 /y :
Z/Z
Ms (≤ Z) 1 − e−Z/y 1 − u0 0
= = . (255)
Ms,0 1 − u0 1 − u0
However, the model does not agree well with the data
because the model is incomplete (infalls, variations in
the IMF, etc.).
• Galaxies whose total luminosity is dominated by radiation not produced in stars. Stars
produce near-UV, optical, and near-IR light in blackbodies. Other sources may emit
radio or X-ray light.
• The energy generation is associated with a point-like source at the nucleus of the galaxy
(∼ black hole with mass 106 − 109 M ).
AGN types:
• Radio galaxies:
56
3. MODELLING GALAXIES
– energized by jets
(particle acceleration Ee ∼ 1012 eV
– highly variable
– extremely luminous
– highly polarized
57
3. MODELLING GALAXIES
– spiral galaxies
M• = 106 M → Rs = 10−7 pc
M• = 107 M → Rs = 10−6 pc (259)
M• = 109 M → Rs = 10−4 pc
We can consider various possible energy sources for the observed variability:
• Stars: N∗ = 3 × 108 O-type stars in the central region to get the necessary luminosity
(O stars have luminosity of ∼ 105.5 L ), but this leads to a stellar density that is too
high and would be unstable.
• Supernovae: the energy of a supernova is ESN ∼ 1052 erg, so we would need 1010 super-
novae within 10−3 pc in 107 years. This would require producing 1010 stars continually,
which has the same problem as the source being stars (too dense and unstable).
Spanish dancer galaxy : NASA, ESA, Hubble; Processing & Copyright: Leo Shatz © 58
Leo Shatz. All rights reserved. This content is excluded from our Creative Commons
license. For more information, see https://ocw.mit.edu/help/faq-fair-use
3. MODELLING GALAXIES
The Eddington luminosity is the maximum possible AGN luminosity, which is reached when
the radiation pressure exceeds the gravitational acceleration per area. This comes from
processes like Thomson scattering. The radiation pressure is given by
E hν
Pγ = = . (261)
c c
We can write the momentum per time (equivalent to force) as L/c, so the pressure is force
per area
L
Ptotal = . (262)
4πr2 c
We find where the radiative force on a fully ionized plasma (i.e. the force on an e− ) exceeds
the gravitational force on a proton for a black hole of mass M• :
Frad > Fgrav
L GM• mp (263)
σ T >
4πr2 c r2
where σT is the Thomson cross section, so the radiative force on an electron is Frad = σT Ptotal .
This gives the Eddington luminosity
4πcGM• mp M•
Ledd = = 1.3 × 1038 erg/s . (264)
σT M
To achieve AGN luminosities, the SMBH must be massive enough to to be blown apart.
This leads to the Eddington accretion rate, which is the maximal possible accretion rate
possible for an accretion disk. This is reached for when Lacc exceeds Ledd :
Lacc > Ledd
1 M• (265)
ṁc2 > 1.3 × 1038 erg/s
16 M
so ṁedd occurs when the two are equal:
M• M
ṁedd = 5 × 10−10 . (266)
M yrs
59
3. MODELLING GALAXIES
An accretion disk is formed when gas spirals in from large distances until the innermost stable
orbit (ISCO). Viscous processes in the disk lead to heating closed
to temperatures
box
T ∼ 108 K. This
Giant Branch
is highly efficient in releasing energy.
log(L) f(≤z)
Zero Age
Unified model of AGN:
Main Sequence data
Different AGN types are manifestations of the same phenomenon:
• Broad line region (BLR) formed f from clouds of thick gas within ∼ 0.1 − 1 pc (velocities
are faster near the black hole, v ∼ 104 km/s
• Narrow line region (NLR) formed from clouds of thin gas within ∼few pc (farther away
from the black hole, v ∼ 100 − 1000 km/s. Slower velocities leads to less broadening
of lines scattered in the clouds, hence narrow line region).
BLR
dust
disk
torus NLR
jet
The observed manifestation depends on the viewing angle and the accretion rate. For exam-
ple, BL Lac objects have a line of sight directly down the jet and a high accretion rate.
60
Part II
61
1. COSMOLOGY
1 Cosmology
Cosmology is the study of dynamics of the entire Universe as a single dynamical system.
• The Universe is isotropic: it looks the same for all observers on large scales.
This implies that the space-time metric is the same everywhere, which generates symmetries
and simplifies the solutions to general relativity equations.
Hubble Law:
We observe that
~v = H0~r (267)
where H0 ∼ 70 km/s/Mpc refers to the present-day Hubble factor. The specific form of this
law can be derived from the cosmological principles:
• Linearity (follows from isotropy):
Suppose ~v = f (~r).
Then from Observer A’s perspective
62
1. COSMOLOGY
Observer a
4
B M (< a) = πa3 ρ
3 (270)
GM (< a) 4πGρ 1
% ⇒ä = − 2
= − 2 · a3
a a 3
4πGρ0 a30 d
d 1 2 1
ȧ = − − . (272)
dt 2 3 dt a
R
Integrate dt on both sides:
1 2 4πGρ0 a30 1
ȧ = + κ̃ . (273)
2 3 a
Here, κ̃ is the integration constant, and we can use the density expression ρ0 a30 = ρa3 to
simplify our equation:
1 2 4πG 2
ȧ = ρa + κ̃ . (274)
2 3
So the dynamics is given by 2
ȧ 8πG k̃
= ρ+ 2 . (275)
a 3 a
This is the Friedmann equation for Λ = 0. With a cosmological constant Λ, we have
2
ȧ 8πG k̃ Λ
= ρ+ 2 + . (276)
a 3 a 3
63
Plot © Geek3, CC BY-SA 4.0, Wikimedia Commons. All rights reserved.
This content is excluded from our Creative Commons license. For more
1. COSMOLOGY information, see https://ocw.mit.edu/help/faq-fair-use/
Since volume grows with the length cubed and the total mass in the universe is constant,
the matter density is proportional to a−3 . The radiation density also decreases due to the
increasing volume but also decreases as the wavelengths are stretched, so radiation density
is proportional to a−4 . The dark energy Λ is constant.
The plot above shows how the scale factor grows with time for several different types of
universes. Our current understanding of our universe is that it is described by the ΛCDM
model, where roughly 30% of the energy budget is matter, 70% is dark energy, and there
is a very small amount of radiation and no curvature. If there were positive or negative
curvature, we would get an open ore closed universe. There are also several toy universes
that are often useful to think about. A flat, dark energy-only universe is the de Sitter model
and a flat, matter-only universe is the Einstein-de Sitter model An empty universe has only
a curvature term, and is an open universe.
Therefore, from the Friedmann equation, we can derive different regimes of the Universe:
• radiation regime
ȧ2 ∝ a−2
1 1 1
ȧ ∝ a−1 =⇒ a ∝ t2 =⇒ t0 =
ȧ 1 2 H0
ada ∝ dt H(t) = a
= 2t
64
1. COSMOLOGY
• matter regime
ȧ2 ∝ a−1
− 21
2 2 1
ȧ ∝ a =⇒ a ∝ t3 =⇒ t0 =
√ ȧ 2 3 H0
ada ∝ dt H(t) = a
= 3t
• curvature regime
ȧ2 ∝ constant
a∝t 1
ȧ ∝ constant =⇒ ȧ 1 =⇒ t0 =
H(t) = = H0
da ∝ dt a t
• Λ regime
2Λ
ȧ2 ∝ aq3
√Λ
ȧ ∝ a Λ3 =⇒ a∝e 3
t =⇒ exponential growth
q
da
a
∝ Λ3 dt
65
1. COSMOLOGY
v1
First we need to specify gµν andObserver
Tµν . a
r2,
B
Observer v2
Metrics:
The ,A
r1-r2space-time interval is
v1-v2 %
r1, ds2 = gµν dxµ dxν . (278)
v1
r2, Some examples of spatial metrics:
Observer a
v2 B
er • 2D-flat space in Cartesian coordinates:
%
2
1 0 dx
ds = dx dy = dx2 + dy 2 (279)
0 1 dy
r1-r2,
v1-v2 • 2D-flat space polar coordinates:
Observer a
B
2
1 0 dr
% ds = dr dθ 2 = dr2 + r2 dθ2 (280)
0 r dθ
• 2D-curved space:
2
R2 0 dθ
ds = dθ dϕ 2 2 = R2 (dθ2 + sin2 θdϕ) (281)
0 R sin θ dϕ
66
1. COSMOLOGY
Robertson-Walker metric:
The metric form follows from homogeneity and isotropy, and the field equations give us the
time evolution:
ds2 = −c2 dt2 + a(t) dχ2 + fk2 (χ) dθ2 + sin2 θdϕ2
(287)
and −1/2
k
sin (k 1/2 χ), closed k > 0
fk (χ) = χ, flat k = 0 (288)
−1/2
|k| sinh (|k|1/2 χ), open k < 0
1
with the units for k: [k] = L2
.
T00 ∼ T0j ∼
= energy density = energy flux
Tµν = (289)
Tj0 ∼
= momentum density Tik ∼
= stress tensor
Tii ∼
= pressure Tik ∼
= shear (290)
To evaluate Gµν , we take the derivative of the metric. We then plug this into the Einstein field
equations, which gives two independent equations. This leads to the Friedmann equations:
2
ȧ 8πG kc2 Λc2
= ρ− 2 +
a 3 a 3
. (293)
Λc2
ä 4πG 3p
=− ρ+ 2 +
a 3 c 3
For relativistic bosons and fermions p = ρc2 /3, and for non-relativistic particles p = 0.
3H 2 (t)
ρcrit = (294)
8πG
67
1. COSMOLOGY
3H02
ρcrit,0 = ≈ 1.8 × 10−29 h2 g/cm3 . (295)
8πG
For a sphere with radius a filled with the critical density, the gravitational potential is equal
to the specific kinetic energy:
G 43 πρcrit a3 ȧ2
= . (296)
a 2
This is the limiting case between an open and closed universe and leads to eternal expansion.
ρm (t)
Ωm (t) =
ρcrit (t)
ρr (t)
Ωr (t) =
ρcrit (t)
kc2 (297)
Ωk (t) = − 2
H
Λc2 ρΛ (t) Λc2
ΩΛ (t) = = , ρ Λ (t) =
3H 2 ρcrit (t) 8πG
Ω(t) = Ωm (t) + Ωr (t)
We also often consider the baryon density parameter Ωb separately from the total matter
density, so the total matter density is the sum of the baryon and dark matter densities
Ωm = Ωdm + Ωb .
68
1. COSMOLOGY
Our current measurements of these values are (from the Planck 2018 results)
Ωm,0 = 0.315 ± 0.007
Ωdm,0 = 0.264 ± 0.003
Ωb,0 = 0.0493 ± 0.0003 (300)
Ωk,0 = 0.0007 ± 0.0019
ΩΛ,0 = 0.6847 ± 0.0073
with H0 = 67.4 ± 0.5 km/s/Mpc. The radiation parameter Ωr,0 can be derived from the
measured temperature of the CMB and relating the photon and neutrino density to get
Ωr,0 ≈ 10−4 . We discuss how to obtain these values from observations in Part III.
So we find:
H 2 (a) = H0 2 E 2 (a)
(302)
E 2 (a) = Ωr,0 a−4 + Ωm,0 a−3 + Ωk,0 a−2 + ΩΛ,0
which is a useful form of the Friedmann equation.
Notes:
• Radiation dominates in early times, then matter, then the cosmological constant.
• Matter-Λ equality occurs when ΩΛ = Ωm :
Ωm,0
ΩΛ,0 = 3
a (303)
1
=⇒a ≈ , z ≈ 1.3 (z ≈ 1 is 6 − 7 Gyr after the Big Bang) .
2.3
• Matter-radiation equality occurs when Ωr = Ωm :
Ωr,0 a−4 = Ωm,0 a−3 (304)
1
=⇒a ≈ , z = 3700 . (305)
3700
• Observationally, Ωk,0 ≈ 0.
69
r1, v1-v2
v1
r2, Observer a
1. COSMOLOGY v2 B
Observer
A
1.C Observational cosmology %
Redshift:
Redshift z is defined by the difference in observed wavelength and emitted wavelength of
light:
λobs λobs − λem
=1+ ≡1+z (306)
λem λem
In cosmology, this is due to the expansion of space. Light travels from the source at
(tem , aem , zem ) to the observer at (tobs , aobs , zobs ).
λ
λobs 'tobs = cobs λem 'tem = λcem
B A B A
Observer at &obs Source at &em
The spatial hypersurface can shrink or expand depending on a(t), so λobs is not necessarily
equal to λem . Photons always travel along the shortest path, so for light we have:
70
1. COSMOLOGY
energy of photons
Note: the change of luminosity L = time
is affected “twice" by expansion since
Distance measures:
Question: what is the distance between a source at (z, t, a) and an observer?
In static Euclidean space, we can measure a unique distance in different ways. For a source
with luminosity L and size l, we have:
L
• luminosity distance DL : F = 4πDL2
l
• angular diameter distance DA : ϕ = DA
Comoving distance:
Comoving coordinates move with space as it expands.
Z tobs
cdt
χ(tem , tobs ) =
tem a(t)
(for light, ds = 0 ⇒ adχ = cdt) (314)
Z aobs Z aobs
obs da c da
= χem = c =
aem ȧa H0 aem a2 E(a)
The comoving distance is not measurable through observations, but it is useful theoretically.
71
1. COSMOLOGY
We also define a function that depends on the curvature k of space that is helpful in writing
the metrics: √
1
√ sin χ k , k>0
k
fk (χ) = χ, k=0 (315)
1
p
p sinh χ |k| ,
k<0
|k|
Angular diameter distance:
The angular diameter distance DA is defined such that
l
ϕ= (316)
DA
for an object that has angular size ϕ. Then the endpoints of l have the same (χ, θ, t):
l = aem fk χobs
em ϕ
1 (317)
fk χobs
= em ϕ
1+z
observer
( em &em
&obs l
obs
observer
emission emission
so
1
fk χobs
DA = em . (318)
1+z
Luminosity distance:
The luminosity distance DL is defined such that
L
F = . (319)
4πDL2
observer
m &em
obs
Then the observed surface, for aobs = 1, is
s
l
72
1. COSMOLOGY
so
DL = (1 + z)fk χobs
em . (322)
This also gives us the relation
1
DA = DL (323)
(1 + z)2
so DA ≈ DL if z 1.
Notes:
• The simplest Einstein-de Sitter case is:
ΩΛ,0 = Ωκ,0 = Ωr,0 = 0, Ωm,0 = 1
2c 1 1
⇒ DA = 1− √
H0 1 + z 1+z (324)
2c 1
DL = (1 + z) 1 − √
H0 1+z
For z 1:
2c 1
DA ≈ DL ≈ 1− 1− z (325)
H0 2
c
= z (326)
H0
Furthermore:
Z aobs Z 1
c da c da
χobs
em = 2
= (327)
H0 aem a E(a) H0 aem a2 E(a)
(328)
For E(a) ≈ 1 and a ≈ 1 − z:
Z 1
c 1
χobs
em ≈ da (329)
H0 (1 − z)2 aem
c 1
= (1 − a) (330)
H0 (1 − z)2
c 1
≈ z (331)
H0 (1 − z)2
c
≈ z (332)
H0
So we get DA = DL = χobs
em for z 1, i.e. agreement for low z.
• The general flat case is: k = 0 ⇒ fk (χ) = χ.
1
DA = χobs
1 + z em Z
aobs
c 1 da
= 2 (333)
H0 1 + z aem a E(a)
Z 1
c 1 da
= 2
H0 1 + z aem a E(a)
73
1. COSMOLOGY
1 1 2
Since a = 1+z
and da = − 1+z dz, then
Z z
c 1 1
DA = [Ωm,0 (1 + z)3 + Ωr,0 (1 + z)4 + ΩΛ,0 ]− 2 dz (334)
H0 1 + z 0
109
DA
DL
108
To the left is a plot com-
107 paring the angular diame-
Distance [Mpc]
102
In general:
V = dAdr (337)
74
1. COSMOLOGY
c fκ2 (χ)
dVχ = dzdΩ (343)
H0 E(z)
dχ
= fk2 (χ)rdΩ dz . (344)
dz
1.D Inflation
So far, dynamics have been described by the Friedmann equations with some mass-energy
content of the Universe: Ωm , Ωr , Ωk , ΩΛ . Is this sufficient to explain all data?
Problems:
• Horizon problem:
ρr ∝ (1 + z)4 and ρr ∝ T 4 ⇒ T ∝ (1 + z) (345)
The Universe cools and at some zrecomb , it consists of neutral hydrogen atoms (recom-
bination). We get the balancing equation
H+ + e H0 + χ (346)
75
1. COSMOLOGY
After this time, photons can escape or free stream, and we can observe them as the
Cosmic Microwave Background. The CMB is very uniform: ∆T T
≤ 10−5 (note that
CMB maps are typically logarithmic).
76
1. COSMOLOGY
1 1
DA = fκ (χobs
em ) = χobs
em (for flat universe)
1+z Z z 1 + z
c 1 − 1
Ωm,0 (1 + z)3 2 dz
=
1 + z H0 0
!z
2c 1
= 1 −√
H0 (1 + z)Ωm 0 2 1+z
0 (355)
2c 1
= 1 1− √
H0 (1 + z) Ωm0 2 1+z
| {z } | {z }
≈1,z1
≈z,z1
2c 1
≈
H0 Ω 12 z
m0
This gives us the angular size of the horizon at recombination:
r
1
ϕhorizon,recomb ≈ ∼ 1.7◦ (356)
zrecomb
Or more generally:
ϕhorizon,recomb ≈ 1.7◦
p
Ωm,0 . (357)
This is much smaller than the full sky, so how can the CMB be so uniform?
• Flatness problem:
At high z, Λ is irrelevant in the Friedmann equations, so:
kc2 kc2
2 8πG 2
H (a) = ρ − 2 = H (a) Ω(a) − 2 2 , ρ = ρm + ρr (358)
3 a a H (a)
kc2
|Ω(a) − 1| = . (359)
a2 H 2 (a)
Since a ∝ t2/3 in matter dominated times and a ∝ t1/2 during radiation dominated
times, we have: (
t, radiation dominated
|Ω(t) − 1| ∝ 2/3 (360)
t , matter dominated
Thus, any small deviation Ω(tearly ) 6= 1 at early times quickly blows up! Ω(tearly ) must
therefore be very close to 1, which leads to a “fine-tuning problem."
• Monopole problem:
General unified theories predict many magnetic monopoles, but this is not observed.
The number density must decrease.
77
1. COSMOLOGY
• Flatness problem:
2
kc
If a2 H 2 (a) decreased with time for a short period, then Ω(a)
observer
emission emission
• Horizon problem:
kc2 c
If a2 H 2 (a) shrinks, then χ ∝ aH(a)
also shrinks.
comoving V())
horizon before slow roll
causally inflation
connected region
⇒ canVexplain
0 smoothness within the observable universe.
comoving
horizon after reheating
1
So decreasing aH(a)
seems to solve two problems! The conditions for a shrinking comoving
)0 )
horizon:
d c
<0
dt aH
d c
<0
dt ȧ (361)
cä
− 2 <0
ȧ
⇒ ä > 0
We need some period of accelerated expansion. We can look at the second Friedmann
equation (e.g. for acceleration):
ä 4πG 3p Λc2
=− (ρ + 2 ) +
a 3 c 3
(at early times, Λ = 0)
4πG 3p
=− (ρ + 2 ) (362)
3 c
2
ρc
⇒p<− ← we need sufficiently negative pressure
3
p 1
⇒ 2 <−
ρc 3
This also solves the monopole and seed problem! Rapid expansion would decrease the density
of monopoles and blow up tiny perturbations. All problems are then solved.
78
1. COSMOLOGY
2
Note that Λc3 actually corresponds to a negative pressure term. To see this more clearly, we
combine both Friedmann equations to derive the energy conservation equation:
d d
(ρc2 a3 ) + p (a3 ) = 0
dt dt (363)
p
⇒ ρ̇ = −3H(a)(ρ + 2 )
c
Inflation:
Λ has all the features we want, but it:
• acts too late
• is constant, i.e. even if it acted early enough, it would not stop inflation!
How do we get all this in the early universe? We look at a homogeneous scalar field (inflation):
1
L = ∂µ φ∂ µ φ − V (φ) (367)
2
which leads to the energy-momentum tensor:
1
T00 = ρc2 = φ̇2 + V (φ)
Tµν = ∂µ φ∂ν φ − gµν L ⇒ 2 (368)
1 2
Tii = p = φ̇ − V (φ)
2
To get w < −1/3, we require:
1 1 1
φ̇ − V (φ) < − φ̇ + V (φ)
2 3 2
ρc2 (369)
p<−
3
2
⇒ φ̇ < V (φ)
79
1. COSMOLOGY
i.e. the field must be moving slowly during inflation. Thus, the potential should be flat and
have a minimum to stop inflation. Furthermore:
2 8πG 1 2
Friedmann equation : H = φ̇ + V (φ)
3 2
p
Energy conservation : ρ̇ = −3H(a) ρ + 2
c
2 dV
ρ̇c = φ̇φ̈ + φ̇
dφ
1 2 p 1
with ρ = φ̇ + V (φ) and 2 = φ̇ − V (φ)
2 c 2
dV 1 2 1 2
⇒φ̇φ̈ + φ̇ = −3H(a) φ̇ + V (φ) − 3H(a) φ̇ − V (φ)
dφ 2 2
dV
⇒φ̈ + = −3H(a)φ̇
dφ
dV
⇒ φ̈ + 3H(a) φ̇ = −
| {z } dφ
Hubble drag
(370)
and we get the field evolution equation. In a static universe, H = 0, and there is no Hubble
drag. dV
dφ
is how fast energy is extracted from inflation.
As long as these conditions are valid, inflation will go on. The slow roll potential is:
80
emission emission
1. COSMOLOGY
ng V())
efore
usally slow roll
ed region V0
During reheating, the in-
ing inflation
flation field decays through
after
reheating coupling to ordinary matter
(“reheat universe").
)0 )
(minimum)
Since
8πG 8πG
H2 = V (φ) ≈ V0 (375)
3 3
during inflation, large values of φ0 and V0 lead to more inflation (longer slow roll).
• metric (geometry)
Emerging story
a) t = 0: Big Bang
b) t ∼ 10−34 s: inflation
c) T decreases as T ∝ (1 + z)
e) z ≈ 1100: recombination
81
2. STRUCTURE FORMATION
The first five stages here are optically thick to photons, while later is optically thin and
potentially observable.
2 Structure formation
So far, we have assumed a uniform cosmology. We now add perturbations to study the
growth of structure.
Basic equations:
∂ρ ~ p
continuity equation : + ∇ · (ρ + 2 )~v = 0
∂t c
~
momentum equation :
∂~v ~ v = − ∇pp + ∇φ
+ (~v · ∇)~ ~ (377)
∂t ρ + c2
~ 2 φ = 4πG(ρ + 3p )
Poisson’s equation: ∇
c2
82
2. STRUCTURE FORMATION
Notes:
• non-relativistic
– Dark matter follows the collisionless Boltzmann equation; 1st /2nd equations only
hold for moments (Jeans equation).
– For dark matter, there is no well-defined velocity field ~v (~x, t) due to multistream.
~v (~x, t) is just an average.
– Nevertheless, it recovers the correct growth rate for large scales > λJ when pres-
sure can be neglected.
• relativistic
This leads to the perturbation equation where some small perturbation δ evolves in a smooth
background density ρ̄:
∆ρ ρ − ρ̄
δ= = . (378)
ρ̄ ρ̄
Perturbation equations:
!
~ 2δ
v2∇
non-relativistic: δ̈ + 2H δ̇ = 4πGρ̄δ + s 2
a
! (379)
32 ~ 2δ
v2∇
relativistic: δ̈ + 2H δ̇ = πGρ̄δ + s 2
3 a
where
δ = δ(~x, t)
~x : comoving coordinates
~r : a~x physical coordinates
∇~ = ∂
∂~x (380)
cs , non-relativistic baryons
vs = σ, non-relativistic dark matter
c
√ ,
relativistic radiation
3
Fourier representation:
83
2. STRUCTURE FORMATION
and we get
vs2 k 2
¨ ˙
non-relativistic: δ̂ + 2H δ̂ = δ̂ 4πGρ̄ − 2
a
(382)
vs2 k 2
¨ ˙ 32
relativistic: δ̂ + 2H δ̂ = δ̂ πGρ̄ − 2 .
3 a
Growing modes:
For a static background, H = 0 and:
¨
δ̂ + w02 δ̂ = 0 (383)
with
v2k2
s 2 − 4πGρ̄, non-relativistic
w02 = a (384)
v 2 k 2 32
s − πGρ̄,
relativistic
a2 3
For physical wave number k̃ = k/a, we get oscillation for:
√
2 πGρ̄
vs , non-relativistic
k̃ ≥ k̃J = q 32 √ (385)
πGρ̄
3
, relativistic
vs
2π
and growth (no oscillations) for modes with lengths l = k̃
greater than λJ :
2π vs
l ≥ λJ = ∝√ (386)
k̃J πGρ̄
where k̃ = k/a is in physical units. We can make this more general for H 6= 0 and neglecting
the pressure terms for l ≥ λJ and we get:
¨ ˙
δ̂ + 2H δ̂ = 4πGρ̄δ̂, non-relativistic
¨ ˙ 32 (387)
δ̂ + 2H δ̂ = πGρ̄δ̂, relativistic
3
Now for Ω = 1, we have the critical density as the background density for the radiation and
matter dominated regime:
3H 2
ρ̄ = ρcrit = (388)
8πG
84
2. STRUCTURE FORMATION
such that
¨ ˙ 3
δ̂ + 2H δ̂ = H 2 δ̂, matter dominated Ω = 1
2 (389)
¨ ˙
δ̂ + 2H δ̂ = 4H 2 δ̂, radiation dominated Ω = 1
Now:
2
ȧ
, matter dominated
H = = 3t (390)
a 1 , radiation dominated
2t
We now assume δ̂(~k, t) ∝ tn . Then:
n 2
n2 + − = 0, matter dominated
3 3 (391)
n2 − 1 = 0, radiation dominated
2
n = −1, , matter dominated
⇒ 3 (392)
n = −1, +1, radiation dominated
that correspond to negative decaying modes, which are unimportant since the perturbations
vanish, and positive growing modes. This gives us:
(
a, matter dominated
δ̂ ∝ (393)
a2 , radiation dominated
In general, we write δ = Dδ0 or δ̂ = Dδ̂0 where D is the growth factor such that D(z = 0) = 1.
• Before recombination
∂p 1
c2s = , p = ρc2 , (radiation)
∂ρ 3
c (394)
⇒cs = √ .
3
• After recombination
∂p ρ
c2s = , p= , T = 2.71 K(1 + z), (ideal gas)
∂ρ mkT
r (395)
kT
⇒cs = ≈ 5 km/s
m
85
2. STRUCTURE FORMATION
• So
z 1100 : MJ ≥ 1016 M
(396)
z < 1100 : MJ . 105 M
since after recombination, the photon pressure support is removed.
• Structure can only form after z ∼ 1000.
• if structure can only grow from z ∼ 1000, δ will be amplified by ∼ 103 (since matter
dominated growth is ∝ a). BUT: the CMB has δ ∼ 10−5 and 10−5 × 103 ∼ 10−2 today,
which is much less than what we observe in the low redshift universe. This theory of
structure growth is not sufficient.
• Solution: dark matter must have clumped before and baryons fall into dark matter
wells.
Cold dark matter:
• CDM is very cold, so it has a tiny velocity dispersion σ. This means that MJ is tiny
and collapse on all scales is possible.
• CDM does not interact with radiation, so it can grow before recombination.
Cold dark matter is needed to make structure formation work!
86
2. STRUCTURE FORMATION comoving V())
horizon before
where we used causally slow roll
2
8πGρ 3H
H 2 =connected (forregion
ρ = ρcrit = )V0
3 8πG
comoving
1 inflation
, radiation (400)
horizon
2 after
= 4t re
4 ,
matter
at2
lhorizon < λJ : as soon as mode length l enters the horizon, it will oscillate! So before
recombination, perturbations can grow if l > lhorizon , otherwise they will oscillate.
(mi
log(t)
The horizon and Jeans mass grow as we have seen before: MJ ∼ 1016 M at z ∼ 1000, so all
modes smaller than 1016 M entered the horizon before recombination and therefore start to
oscillate and stop growing. mode enters recombination
matter horizon
radiation
There is also another problem for those modes: Silk damping! Before decoupling, photons
dominated dominated
do not free stream because of Thomson scattering off free electrons. The mean free path
log(')
gets large towards recombination. So: 2/3 ∝t , a
16
- M < MJ ∼ 10 M perturbations oscillate due to photon pressure.
oscillations
∝t, a2 ∝t2/3, a
- Photons can diffuse out of potential wells and take baryons with them (electrons
through Thomson scattering and protons through Coulomb interactions), which erases
perturbations.
The net effect is that all perturbations ∼ 1012 M (Silk mass) are damped and erased!
log(t)
Cold dark matter:
Cold dark matter has essentially zero Jeans mass, so all modes can already grow. However,
for subhorizon modes in the radiation dominated epoch, δ ∼constant (stagnation). Because
87
)0 )
(minimum)
log(') ∝a ∝a
stagnation
∝a2 ∝a2
Cold dark matter is then the main driver of structure formation since it there is time for CDM
perturbations to grow large enough. Without CDM, structure formation is not possible.
variance:
σ 2 = δ 2 − hδi2 = δ 2 > 0
d3 k (404)
Z Z
2 3 2
δ = d xδ (~x) = |δ̂(~k)|2
(2π)3
88
2. STRUCTURE FORMATION
If we assume homogeneity and isotropy, ~k → k = |~k| and d3~k = 4πk 2 dk. Then we get:
Z
1
2
σ = 2 |δ̂(k)|2 k 2 dk
2π
Z
1 (405)
=: 2 P (k)k 2 dk
2π
with P (k) = |δ̂(k)|2
Notes:
• P (k) and σ are functions of time since δ̂(k) grows (σ = Dσ0 ).
• The initial power spectrum is the primordial power spectrum set at the end of inflation.
The general form is
P (k) = Ak n (406)
which is a power law and is scale-free. According to predictions from inflation, n ≈ 1.
Measuring P (k) and galaxy clustering:
If we assume galaxies trace the mass perturbations, what is the probability dP that we find
two galaxies in volumes dV1 and dV2 at a distance r from each other?
dP = n0 (1 + δ(~x))dV1 · n0 (1 + δ(~x + ~r))dV2
= n20 (1 + δ(~x) + δ(~x + ~r) +δ(~x)δ(~x + ~r))dV1 dV2
|{z} | {z } (407)
=0 =0
= n20 (1 + ξ(r))dV1 dV2
where δ(~x) and δ(~x + ~r) are zero on average and ~r → r due to isotropy. ξ is the two-point
correlation function and is related to P (k):
Z
ξ(r) = d3~xδ(~x)δ(~x + ~r)
d3 k ~ −i~k·~x d3 k 0 ~ 0 −i~k0 ·(~x+~r)
Z Z Z
3
= d ~x δ̂( k)e δ̂(k )e
(2π)3 (2π)3
| {z }
d3 k 0 0
δ̂(~k0 )e+i~k ·(~x+~r)
R
(δ real)
(2π)3
2 Z Z Z
1 ~ ~0 ~
= d3~x d3 k d3 k 0 δ̂(~k)δ̂(~k 0 )e−i(k−k )·~x e−ik~r
(2π)3
Z
1 ~ ~0
using d3 xei(k−k )·~x = δ(~k − ~k 0 ) (408)
(2π)3
Z
1 ~
= |δ̂(~k)|2 eik·~r d3 k
(2π)3
Z
1 ~
= |δ̂(k)|2 eik·~r d3 k where ~k → k from isotropy
(2π)3
Z
1 ~
= P (k)eik·~r d3 k
(2π)3
Z
1 ~
⇒ ξ(r) = P (k)eik·~r d3 k .
(2π)3
89
2. STRUCTURE FORMATION
Observationally: −1.8
r
ξ(r) ≈ (409)
r0
with r0 ≈ 5 h−1 Mpc for galaxies. Different objects have a different r0 , and more massive
objects are more clustered, e.g. the cluster-cluster correlation function differs from the
galaxy-galaxy correlation function: ξcc ≈ 20ξgg .
The horizon mass, i.e. the mass within the horizon, is:
Mh ∝ ρm rh3
(413)
ρm ∝ (1 + z)3
and ( 2
a = (1 + z)−2 , radiation dominated
rh ∝
a3/2 = (1 + z)−3/2 , matter dominated
( (414)
(1 + zh )−3 , radiation dominated
⇒ Mh ∝
(1 + zh )−3/2 , matter dominated
90
2. STRUCTURE FORMATION
σ grows:
σ∝δ
( 2
a = (1 + z)−2 , radiation dominated (415)
σ∝
a3/2 = (1 + z)−1 , matter dominated
We now find σ of the horizon mass, i.e. the fluctuation strength once this mode enters.
where zh is the redshift once mass M is within the horizon, and zp is the redshift at
the end of inflation. Then
1
σ(zp ) ∝ M − 6 (n+3)
(417)
Mh = M ∝ (1 + zh )−3 ⇒ (1 + zh )−2 ∝ M 2/3
so we find:
1 1 n−4
σ(zh ) ∝ M − 6 (n+3) M 2/3 = M −( 2 + 6
)
(418)
Now, we do not want “special" modes, so σ(zh ) should not depend on n! We get n ≈ 1
according to the Harrison-Zel’dovich spectrum.
where b is the bias of the galaxy clustering compared to the mass fluctuations. Observation-
ally, σ8,gal ≈ 1. From WMAP and SDSS, we have:
n = 0.953 ± 0.016
(421)
σ8 = 0.756 ± 0.035
Transfer function:
We found that modes entering the horizon during the radiation dominated phase do not grow
91
2. STRUCTURE FORMATION
(stagnation). The primordial power spectrum is therefore modified by the transfer function:
P0 (k) = (Ak)T 2 (k),
1
1, L0
(422)
T (K) ≈ k
, 1 L0
1
k2 k
where L0 is the comoving horizon at zequality .
Halos:
We can simplify:
−3/2
τ = Hta t (with Hta = H0 ata )
dx 1 ȧ H
⇒ x0 = = = x = x−1/2
dτ Hta ata Hta (424)
H H0 a−3/2 a−3/2
(using = −3/2
= −3/2
= x−3/2 for the final equality)
Hta H0 ata ata
So
x0 = x−1/2 (425)
92
2. STRUCTURE FORMATION
ξ
y 00 = − (428)
2y 2
2
x0 = x−1/2 ⇒ τ = x3/2 (431)
3
So
3
x= τ 2/3
2
2/3 (432)
dx 2 3
= τ −1/3 = x−1/2
dt 3 2
We also have r
0
p 1
y =± ξ −1 (433)
y
using the first boundary condition. We also use the + before turnaround and the − after.
93
2. STRUCTURE FORMATION
Then 1/2 !
dy 0
p 0d 1
= ± ξy −1
dτ dy y
−1/2
p 01 1
= ± ξy −1 (−y −2 )
2 y
−1/2 ! (434)
ξ 1 0 1
= − 2 ±√ y −1
2y ξ y
r −1
ξ 0
p 1
=− 2 y ± ξ −1
2y y
| {z }
=1
Integrating before turnaround and using the second boundary condition gives us an implicit
solution for y:
1 1 p π
τ=√ 2
arcsin(2y − 1) − y − y + . (435)
ξ 2 4
At turnaround:
2
x = 1 = y, τ =
3
2 1 1 π 1 π
⇒ = √ arcsin(1) + = √ (436)
3 ξ 2 | {z } 4 ξ2
π/2
2
3π
⇒ξ=
4
so we get the overdensity parameter ξ.
At collapse:
We assume symmetry, so we get collapse at τ = 43 . Then
2/3 2/3 2/3
3 2/3 3 4
xc = τ = = 41/3 (437)
2 2 3
Collapse parameters:
94
2. STRUCTURE FORMATION
Note here that ξ is the overdensity at turnaround, and we want to find ∆ at collapse.
The background density is proportional to x−3 , and the halo density is proportional to
y −3 . We also know that x = y = 1 at turnaround, where ∆ = ξ. Now
2 3/2 2 3/2 8 3/2 3y
τ= x ⇒ x ≈ y 1+
3 3 9π 10
3/2
x 3 8 3y
⇒ = 1+
y 2 9π 10
3 2 2
x 4 3y (440)
⇒ = 1+
y 3π 10
| {z } | {z }
=1/ξ ≈(1+ 3y
5 )
3
x 3y
⇒∆ = ξ =1+
y 5
linear density contrast (assuming y 1):
3y
δ =∆−1= (441)
5
– The linearly extrapolated density contrast at turnaround is:
ata δ 3y
δta = δ= = (442)
a x 5x
since linear perturbations δ grow like the scale factor. Now
−2/3 2/3
1 3τ 3π 1
= ≈ (443)
x 2 4 y
using the lowest order in y. We can then insert this into δta and get:
2/3
3 3π
δta = ≈ 1.06 (444)
5 4 a
R
– The
P0(k)linearly extrapolated density contrast at collapse is:
2/3
ac 1/3 3 3π
δc = δta = xc δta = 4 δra = ≈ 1.69 (445)
ata 5 2
k0=1/L0 k
So the halo can be considered collapsed when its density contrast expected from lineary
theory has reached δc . If we draw a density field as a function of one-dimensional space,
we can identify which overdensities will collapse at a given time:
collapsing overdensities
'(x) '(x)
'c 'c
time
x x
95
2. STRUCTURE FORMATION
• Nonlinear values:
We now look at the potential energy of a halo:
at turnaround: E = Vta (no kinetic energy)
1
at collapse: E = Tc + Vc = Vc (virial theorem: 2Tc + Vc = 0) (446)
2
1
⇒Vra = E = Vc ⇒ Vc = 2Vta
2
Since potential energy is proportional to 1r and y = 1 at turnaround, we know that
y = 12 at virialization. Then we get the overdensity at this time:
3 1/3 3 2
xc 4 3π
∆V = ξ= 1 ξ = 32ξ = 32 = 18π 2 ≈ 178 (447)
y 2
4
A halo in virial equilibrium is expected to have a mean density of ∼ 178 higher than
the background. This is why masses and radii of halos are often quoted as M200 , which
is the mass enclosed in a sphere of radius R200 with an average density 200 times the
mean or critical density of the Universe.
Analytic derivation:
We consider a halo of mass M . The characteristic length scale is then R(M ) = R:
4π 3
R ρc (z)Ωm (z) = M
3
1/3 (448)
3M
⇒R(M ) =
4πρc (z)Ωm (z)
Halos of mass M are then forming if the smoothed density field δ̄ crosses δc = 1.69:
Z
δ̄(~x) = d3 yδ(~x)WR (|~x − ~y |) (449)
Inflation produces a Gaussian random field, so the probability of finding a smoothed density
contrast δ̄(~x) at a given point in space ~x is:
δ̄ 2 (~
x)
1 −
2σ 2 (z)
p(δ̄(~x), z) = p e R (451)
2πσR2 (z)
96
2. STRUCTURE FORMATION
where σR (z) is the linearly evolved σR : σR (z) = σR D(z) for growth factor D.
The Press-Schechter idea is that the probability of finding the filtered density contrast at
or above the linear density contrast for spherical collapse, δ̄ > δc , is equal to the fraction of
volume filled with halos of mass M :
Z ∞
1 δc
F (M, z) = dδ̄p(δ̄, z) = erfc √ . (452)
δc 2 2σR (z)
∂F (M,z)
The distribution of halos over mass M is simply ∂M
. To calculate this, we need:
However, we need a fudge factor for the mass function to work. We require
Z 1
∂F (M, z)
dM =1 (457)
0 ∂M
∂F
since ∂M is a volume fraction. But we get 12 using ∂F∂M
(M,z)
above! We therefore add a factor
of two: r 2
2 ρc (z)ΩM (z)δc d ln(σR ) − 2σ2 δDc2 (z) 1
N (M, z) = e R . (458)
π σR D(z) dM M
97
2. STRUCTURE FORMATION
See Problem Set 6 for a description of the Extended Press-Schechter formalism that explains
the fudge factor.
98
Part III
99
a
R
P0(k)
1. THE COSMIC MICROWAVE BACKGROUND
So far, we have mostly discussed the late-time evolution of the Universe (except inflation).
We now study the earlykphases.
0=1/L0 k
At z ∼'1000,
c photons decouple from matter (previously coupled'cdue to Thomson scattering).
time matter potential wells, which leads
At that time, the dark matter has already formed dark
to perturbations in baryons. This leads to temperature fluctuations in the CMB δT
T
∼ 10−5
δT −3
(note: without dark matter, we would expect T ∼ 10 ). This leads to anisotropies in the
CMB. x x
Primary anisotropies:
These are anisotropies caused by properties of the CMB.
• Large scales:
• On scales larger than the horizon, baryons follow dark matter, leading to higher tem-
peratures in dark matter wells.
• On scales smaller than the horizon, baryons feel radiation pressure. This leads to
baryonic acoustic oscillations (BAO).
• On very small scales, the imperfect coupling between photons and electrons leads to
diffusion. Fluctuations are smeared out and damped on scales ≤ 50 . This is called Silk
damping.
Secondary anisotropies:
These impact the measurements of the CMB due to effects on photons as the travel from
the CMB to us.
• Thomson scattering of CMB photons: the Universe was reionized by the first stars,
galaxies, and quasars between z ∼ 1000 and z ∼ 6. These photons then experience
Thomson scattering with free electrons as they travel through space. The scattering is
isotropic, so it results in an overall reduction of CMB anisotropies.
100
collapsing overdensities
'(x) '(x)
1. THE COSMIC MICROWAVE BACKGROUND
'c 'c
time
• Integrated Sachs-Wolfe effect: photons experience gravitational potential and time
delays as they travel through structures in the Universe.
x
• Gravitational lensing from structures in the Universe.
•
Sunyaev-Zel’dovich (SZ) effect: CMB photons passing
intensity
through the hot intergalactic medium of galaxies Thom-
son scatter with electrons. This reduces the intensity for
ˠ
lower frequencies and increases the intensity for large
frequencies, resulting in a shift in the spectrum.
frequency
• Large scales: Sachs-Wolfe and Doppler effects roughly compensate each other. The
photons then provide an imprint of the dark matter distribution.
We need to quantify the temperature fluctuations on the sky. We decompose the fluctuations
into spherical harmonics: X
~ =
T (θ) ~
alm Y m (θ) (459)
l
l,m
since Z 2π Z π
dϕ dθ sin θYl1m1 ∗ (θ, ϕ)Yl2m2 (θ, ϕ) = δl1 l2 δm1 m2 . (461)
0 0
We define the power spectrum:
Cl = |alm |2 (462)
averaging over m. We often plot l(l + 1)Cl and define this as the amplitude of fluctuations
◦
on the angular scale θ ∼ πl = 180l . l = 1 is the dipole anisotropy due to the motion of Earth
and l = 2 is the quadrupole anisotropy.
101
k0=1/L0 k
'c universe: ϕhorizon,rec ≈ 1.7◦ . Then for θ 'c1.7◦ , large-scale effects dominate
so for a flat
(Sachs-Wolfe and Doppler) and there are no baryonictimea acoustic oscillations. Then Cl re-
R
flect the 0matter power spectrum P (k) on large scales. For P (k) ∝ k (Harrison-Zel’dovich
P (k)
◦
spectrum), l(l + 1)Cl is approximately constant
x for l 180
1.7◦
≈ 100. x
intensity
√
The baryon-photon fluid has a sound speed of cs ≈ c/ 3. Then the
largest wavelengths such that the wave can have half an oscillation
collapsing overdensities
(compression)
'(x) until zrec
ˠ, trec is: '(x) DM potential
well
baryons
'c c 'c
frequency
λmax = trec cs = trec √ time
. (464)
n=1 3
The sound horizon is √13 times smaller than the horizon. The angular scale is then
x x
1.7◦
θ1 ≈ √ ∼ 1◦
n=2 3 (465)
l1 ≈ 200
intensity
so we can expect the first peak in l(l + 1)Cl there since baryons are compressed. Adiabatic
compression and the Doppler effect lead to temperature fluctuations on that scale.
ˠ
n=3
baryons
The second peak occurs for scales for which one full oscillation is possible and so forth:
frequency
n=1
n=2
n=3
Peaks happen at stationary points of oscillations. We can draw the power spectrum:
102
1. THE COSMIC MICROWAVE BACKGROUND
t1
103
t2
1. THE COSMIC MICROWAVE BACKGROUND
6000
5000
4000
D`T T [µK2 ]
3000
2000
1000
0
600
101 200
300 100
∆D`T T
0 0
-300 -100
-600 -200
` 10
∆D`T T
-10
2 10 30 500 1000 1500 2000 2500
`
Figure 1: [Fig. 57 of Planck Collaboration V. 2020, A&A, 641, A5] Planck 2018 temperature power spectrum. At multipoles
` ≥ 30 we show the frequency-coadded temperature spectrum computed from the Plik cross-half-mission likelihood, with
foreground and other nuisance parameters fixed to a best fit assuming the base-ΛCDM cosmology. In the multipole range
2 ≤ ` ≤ 29, we plot the power-spectrum estimates from the Commander component-separation algorithm, computed over 86% of
the sky (see Sect. 2.1.1). The base-ΛCDM theoretical spectrum best fit to the likelihoods is plotted in light blue in the upper
panel. Residuals with respect to this model are shown in the middle panel. The vertical scale changes at ` = 30, where the
horizontal axis switches from logarithmic to linear. The error bars show ±1σ diagonal uncertainties, including cosmic variance
(approximated as Gaussian) and not including uncertainties in the foreground model at ` ≥ 30. The 1σ region in the middle
panel corresponds to the errors of the unbinned data points (which are in grey). Bottom panel: difference between the 2015
and 2018 coadded high-multipole spectra (green points). The 1σ region corresponds to the binned data errors. The vertical
scale differs from the one of the middle panel. The trend seen for ` < 300 corresponds to the change in the dust correction
model described in Sect. 3.3.2.
Planck Collaboration. Planck 2018 results - V. CMB power spectra and likelihoods. A&A 641 A5 (2020).
License CC-BY.
104
Ωm flat
l Ωk
2. THERMAL HISTORY OF THE UNIVERSE
sound
l(l+1)cl horizon
Ωb
a
Ω2 open
Ω1 > Ω2 leads to t1 < t2 , so there’s notclosed
as much
Ω1
Ωm time to form structures and we get a smaller
flat peak!
l
t1
t2
Note: based on the first peak, we get Ωm and Ωk , so also ΩΛ (assuming flat universe)!
a l(l+1)cl
• Derive Ωb from the second peak:
Ωb is degenerate with ΩΩ
m , so we need the second peak. This can also be derived from
2
Big Bang nucleosynthesis. Ω1 Silk damping
below ~0.5˚
low
baryons high
baryons
With more loading, the mass falls deeper but rebounds to the same position. Thus,
odd peaks are associated with compression—i.e., how deep the baryons fall into the
well. Those peaks get enhanced with more baryons, so the second peak is compressed
compared to the first peak. We can therefore constrain Ωb with the ratio of the two
peaks.
105
2. THERMAL HISTORY OF THE UNIVERSE
We first discuss in some more depth the freeze-out of dark matter, which happens around
10 − 100 GeV. We then briefly discuss the remaining thermal history for temperatures below
∼ 16 GeV (standard model physics). Big Bang nucleosynthesis will be discussed in the next
chapter.
There are two competing timescales: the expansion of the Universe and the mean interaction
timescale. We can rewrite the above equation:
" 2 #
dnc da n c
= −hσvin2c,eq −1
da |{z}
dt nc,eq
ȧ
ȧ (471)
H=
a
" 2 #
a dnc hσvinc,eq nc
=− −1 .
nc,eq da H nc,eq
106
2. THERMAL HISTORY OF THE UNIVERSE
There are two timescales in this equation: τ = 1/H and τcoll = 1/(neq hσvi). Then
" 2 #
a dnc τH nc
=− −1 . (472)
nc,eq da τcoll nc,eq
At redshift zfreeze , we have τcoll ∼ τH , so particles freeze out of equilibrium and the comoving
number density stays fixed.
From observed relic abundances, we get m and σ at the electroweak scale, which is predicted
for WIMPs! This is known as the WIMP miracle. So far, however, nothing has been detected.
107
2. THERMAL HISTORY OF THE UNIVERSE
This wipes out structures on small scales! Since v ≈ c, scales below the horizon are sup-
pressed. But particles slow down due to expansion:
vpec = v − Hd ⇒ v ∝ a−1 (475)
and the particle becomes non-relativistic once mc2 ∼ 3kT (relic time and temperature), so:
−2
mc2
nc,eq 12
nc t ∼ 2 × 10 s
freeze-out 2 keV
−1 (476)
mc2
lh ∼increasing
cth ∼ 60 Mpc relic
<σv> 3 keV
abundance
so hot dark matter erases all structures below lh due to free-streaming.
time
Cold dark 1/T
matter:
Cold dark matter is already non-relativistic at freeze-out, so structures can grow. It still has
some free-streaming scale, but it is much smaller.
primordial
∝k Hot dark matter would not be captured by small
P(k) potential wells, so it needs large potential wells to
∝k-3
for CDM
form structures. This leads to top-down structure
HDM
formation, where large large structures form first
and fragment into smaller structures. Cold dark
matter can form small halos that merge into larger
impact of ones in bottom-up formation.
transfer function T(k)
108
3. BIG BANG NUCLEOSYNTHESIS
Proton/neutron reactions:
After n, p production from the gluon-gluon plasma:
n + νe p + e−
(477)
n + e+ p + ν̄e
with weak interactions mediated by neutrinos and
3/2
mkT mc2
neq = g e− kT . (478)
2π~
Protons and neutrons have gn = gp = 2, so
3/2
nn mn (mm −mp )c2
= e− kT . (479)
np mp
109
3. BIG BANG NUCLEOSYNTHESIS
Deuteron fusion:
We have the strong reaction:
p+n D+γ (482)
with the binding energy of deuteron approximately 2.22 MeV. At the time of neutrino freeze-
out, the temperature is already smaller than 2.22 MeV, but there are so many more photons
than baryons. The high energy tail of photons is still sufficient to destroy deuteron, so we
need kB T 2.22 MeV to efficiently form deuteron! (The deuteron fusion bottleneck means
that 4He fusion afterwards is quick.)
The time delay needed for the temperature to drop below 2.22 MeV causes neutrons to decay
through β-decay before they can fuse to deuteron. Without fusion to deuteron, all neutrons
would be gone!
Note: If we assume the deuteron fusion is instantaneous, what helium/baryon mass fraction
(Y ) would we get?
All neutrons would fuse into 4He :
• think of a group of 12 nucleons: 10p + 2n (5:1 ratio, see above)
• all neutrons fuse into 4He , so we get one 4He atom and eight free protons
•
4 4
⇒Y = = ≈ 0.33 (483)
4+8 12
but we observe 0.24, which is lower due to β-decay.
We now do the precise calculation:
• p+n D + γ never freezes out. It stops once all neutrons are used up. We can use
the Saha equation to find the abundance:
3/2 −3/2
nD gD mD kT 2.22 MeV
= 2
e kT (484)
np nn gp gn mp mn 2π~
where gp = gn = 2 (2 spin configurations) and gD = 3 (3 spin configurations: ↑↑, ↓↓, ↑↓).
We also know mp = mn = mD /2, so:
−3/2
nD mn kT 2.22 MeV
=6 2
e kT (485)
np nn π~
110
3. BIG BANG NUCLEOSYNTHESIS
• Protons always outnumber neutrons, so define the time of deuteron fusion is the time
when half the neutrons have fused into deuteron: nD = nn . Then from the Saha
equation:
−3/2
nD mn kT 2.22 MeV
= 1 = 6np 2
e kT (486)
nn π~
We now want to know when this happens. We can relate np to the temperature to
calculate the corresponding temperature and time. We can relate np to nb , which we
can relate to nγ through the fixed baryon-to-photon ratio η.
6
nb = np + nn = np (487)
5
since nn = 0.2np = 15 np . Then, for a black-body,
3 !
np 5 5 5 kT
= ⇒ np = ηnγ = η 0.24 (488)
nb 6 6 6 ~c
so we get:
3 −3/2
5 kT mn kT 2.22 MeV
1=6 η · 0.24 e kT
6 ~c π~2
3/2
(kT )2 π~2
0.24 · 5 2.22 MeV
= η e kT
1.25 )~c)2 mn kT
3/2
3/2 kT 2.22 MeV
= η |π {z · 1.25} e kT
mn c2
(489)
5.5 · 1.25 = 6.9
3/2
kT 2.22 MeV
≈ 6.9η 2
e kT
mn c
(fiducial η ∼ 5 × 10−10 )
3/2
−9 kT 2.22 MeV
≈ 3.4 × 10 2
e kT
mn c
So we get
T ≈ 8 × 108 K, t ≈ 200 s (490)
for the time of deuteron fusion!
• We lose neutrons through β-decay with a half-life t1/2 = 890 s. After 200 s,
nn
≈ 0.15 < 0.2 (491)
np
111
3. BIG BANG NUCLEOSYNTHESIS
since nHe = nn /2 (every 4He nucleus has 2n) and nH = np − nn (since 4He nucleus has
equal number of protons and neutrons), leaving us with
2(nn /np )
Y = ≈ 0.25 (493)
1 + (nn /np )
Notes:
• Large Ωb leads to larger η, so deuteron can form earlier and there is less neutron decay.
This leads to a larger nn /np , so Y increases with Ωb .
112
Part IV
Selected Topics
113
1. THE LYMAN-α FOREST
So far, we have gone through the basic concepts of the early universe, galaxies, and structure
formation. We have built the basis to discuss some more advanced topics.
The quasar emits at 1216 Å from the hydrogen n = 2 to n = 1 transition. This emission
line is redshifted as it travels through space. The light also passes through neutral hydrogen
clouds, which absorb at 1216 Å, and these absorption lines are also redshifted as the light
continues to travel through space. By the time the light reaches Earth, there is a series
of absorption lines redshifted from 1216 Å, so the absorption lines are observed at different
wavelengths.
By observing quasar spectra passing through the intergalactic medium (IGM), we can use
the Lyman-α forest to probe the density, ionization, temperature, chemistry, and structure
of the IGM. Below are a few examples of quasar spectra. At higher redshift, there are more
opportunities for the light to pass through neutral hydrogen clouds.
114
1. THE LYMAN-α FOREST
Notes:
• Identify lines in general through doublets: µ0II (λ = 2795Å, 2802Å...)
• Lyman-α forest is only visible if it extends to shorter wavelengths than the observed
Lyα emission line at (1 + zem )1216 Å. Photons emitted with 1216Å(1 + zem < λ) <
1216Å will have at some point along the line of sight the right rest frame wavelength
(1216Å) to be absorbed.
– column density NH . 1017 cm−2 gives narrow lines, i.e. the forest
– column density NH & 1017 cm−2 are Lyman-limit systems, i.e. photons with
λ . 912Å = 13.6eV in the rest frame are completely absorbed as the light moves
through the cloud
– column density NH & 1032 cm−2 are damped Ly-α systems, i.e. absorption lines
become very broad.
115
1. THE LYMAN-α FOREST
neutral hydrogen then a quasar spectrum should be totally absorbed. When the redshift is
high enough so that the hydrogen is almost all neutral, the high-wavelength region of the
spectrum will be almost totally absorbed. This is called the Gunn-Peterson trough.
What is observed?
This implies that hydrogen above z ≈ 6 is mostly neutral and mostly ionized below z ≈ 6.
116
1. THE LYMAN-α FOREST
The proper length dl can be related to redshift (taking only the magnitude and ignoring
signs):
da c da c da da
dl = cdt = c = = = (1 + z)c (496)
ȧ a ȧ/a aH H
da
and since a = 1/(1 + z) then dz
= 1/(1 + z)2 so
1 dz cdz
dl = c = (497)
1 + z H(z) (1 + z)H(z)
giving −1
cdz
q
dl = 3
= c H0 (1 + z) ΩM,0 (1 + z) + ΩΛ dz . (498)
(1 + z)H(z)
117
1. THE LYMAN-α FOREST
cdz
dl = p 5 . (499)
H0 ΩM,0 (1 + z) 2
and we can assume that φ, which has a Gaussian shape, is very narrow so is approximately
a delta function Z z 0
c 0 nHI (z ) 0
≈ σ0 1 dz 5 δ(ν(1 + z ) − ν0 )
0
(1 + z ) 2
H0 ΩM,0
2 0
(501)
nHI (z) c 1
= σ0 1 3
H Ω 2 ν0 (1 + z) 2
0 M,0
c
and using λ0 = ν0
, we get
nHI (z)λ0
τν (z) = σ0 1 3
(502)
H0 ΩM,0 (1 + z)
2 2
where σ0 = 10−2 cm2 and z is the redshift when λ(1 + z) = λ0 , i.e. when absorption happens.
This gives the optical depth for one frequency, so is necessary to evaluate at many frequencies
to get the optical depth for different parts of a spectrum.
As an example, we can evaluate τν (z = 3) assuming hydrogen is uniformly distributed and
neutral. Using
There are a couple possible solutions to this discrepancy. It could be that gas isn’t in
intergalactic space. However, we know that this is not the case since we have observed it.
The other option is that the gas isn’t neutral. To bring τLyα (z = 3) down to ∼ 1, we need
to have the neutral hydrogen fraction XHI ≡ nnHI
H
∼ 10−5 .
Ionization:
How does the hydrogen get ionized?
118
1. THE LYMAN-α FOREST
Ionization can occur in hot temperatures. We can estimate the temperature from absorption
line widths:
∆v ∼ 20 km/s
1 2
mv ∼ kT
2 r (504)
2kT
⇒20 km/s ∼
m
⇒T ∼ 30, 000 K ∼ 3eV
which is not enough to ionize hydrogen.
Another option is photoionization. Integrated light from galaxies and quasars emits Γ ∼
10−12 ionizing photons per second. The ionization rate is then ΓnHI , and the ionization
timescale is
nHI
∼ 1012 s ∼ 30, 000 yr (505)
nHI Γ
−0.7
We can compare this with the recombination rate RnHII ne where R = 4.3 × 10−13 10T4 K
which gives the recombination timescale
nHII
∼ 2 × 1017 s ∼ 3 × 106 yr (506)
Rne nHII
so recombination is much slower than ionization and photoionization is plausible.
To establish the predicted ionization fraction, we find equilibrium by setting the ionization
and recombination timescales equal:
119
1. THE LYMAN-α FOREST
120
MIT OpenCourseWare
https://ocw.mit.edu/
8.902 Astrophysics II
Fall 2023
For information about citing these materials or our Terms of Use, visit:
https://ocw.mit.edu/terms.