Cosmology Lecture Notes 2023 (Jonathan Pritchard)
Cosmology Lecture Notes 2023 (Jonathan Pritchard)
1 Introduction:
Homogeneity and Isotropy 1
4 Cosmography 29
4.1 The Age of the Universe . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
4.2 Horizons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
4.3 Distances . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
4.3.1 Luminosity Distance . . . . . . . . . . . . . . . . . . . . . . . . . 33
4.3.2 Angular-Diameter Distance . . . . . . . . . . . . . . . . . . . . . 35
4.3.3 The Extragalactic Distance Ladder . . . . . . . . . . . . . . . . . 35
iii
CONTENTS
7 Big-Bang Nucleosynthesis 57
7.1 Initial Conditions for BBN: neutron-proton freeze-out . . . . . . . . . . . 58
7.2 Helium production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
7.2.1 Dependence upon the cosmological parameters . . . . . . . . . . . 60
7.3 Observations of primordial abundances . . . . . . . . . . . . . . . . . . . 61
8 Interlude 63
8.1 Natural Units . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
8.1.1 The Cosmological Constant problem . . . . . . . . . . . . . . . . 66
8.2 Open Questions in the Big Bang Model . . . . . . . . . . . . . . . . . . . 68
8.2.1 The Flatness Problem . . . . . . . . . . . . . . . . . . . . . . . . 68
8.2.2 The Horizon Problem . . . . . . . . . . . . . . . . . . . . . . . . . 69
8.2.3 The relic particle problem . . . . . . . . . . . . . . . . . . . . . . 70
9 Inflation 73
9.1 Accelerated Expansion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
9.1.1 Acceleration and negative pressure . . . . . . . . . . . . . . . . . 76
9.1.2 The duration of Inflation . . . . . . . . . . . . . . . . . . . . . . . 77
9.2 Inflation via a scalar field . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
9.2.1 Density Perturbations . . . . . . . . . . . . . . . . . . . . . . . . 80
10 Structure Formation 83
10.1 Notation and Preliminaries . . . . . . . . . . . . . . . . . . . . . . . . . . 83
10.2 Spherical Collapse . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
10.3 Linear Perturbations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
10.3.1 Newtonian Theory—non-expanding . . . . . . . . . . . . . . . . . 88
10.3.2 Perturbation theory in an expanding Universe . . . . . . . . . . . 90
10.4 The processed power spectrum of density perturbations . . . . . . . . . . 94
10.4.1 Initial Conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
10.4.2 The transfer function . . . . . . . . . . . . . . . . . . . . . . . . . 96
10.4.3 The effect of baryons: BAOs and the CMB . . . . . . . . . . . . . 99
iv J. R. Pritchard
Preface
These notes were originally produced by Prof. Andrew Jaffe for the Cosmology course as
run between 2008-2012. Further updated by Dr Jonathan Pritchard (2013-2016) and Dr
Subu Mohanty (2017-2021).
Very little of the following is new. Most of it comes, in one form or another, from one
of the following books.
• Rocky Kolb & Mike Turner, The Early Universe
• Andrew Liddle, An Introduction to Modern Cosmology
• Michael Rowan-Robinson, Cosmology
• Peter Schneider, Extragalactic Astronomy and Cosmology: An Introduction
In some cases I explicitly give a reference at the beginning of a chapter or section.
Course structure
The course is structured to cover four main themes in cosmology. Chapters 1-4 establish
the essential language and mathematics needed to talk quantitatively about the physics
of an expanding universe. This lays the foundations for everything that comes later and
enables us to move beyond the level of popular science books to a deeper understanding
of cosmology.
Chapters 5-7 then cover the thermodynamics and physics of a hot big bang. As
the primeval plasma expands and cools the Universe passes through a number of phase
transitions. The number and type of particles produced in these transitions provide one
of the key observational constraints on the physics of the early Universe.
In Chapters 8 and 9, we discuss inflation to address questions of how a big bang might
have occurred and describe some of the traditional problems that an early inflationary
period can solve. Inflation is our best guess at the physics in the first fraction of a
second after the big bang and represents the frontier of our current understanding of the
Universe’s history.
Finally in Chapter 10, we move beyond describing a smooth universe to describe the
irregularities that lead to the formation of structures and galaxies. Modern observations
of the cosmic microwave background and of the positions of millions of galaxies have
provided enormous amounts of information on the distribution of matter in the Universe.
This chapter sketches the theoretical framework for interpreting that data.
v
Chapter 1
Introduction:
Homogeneity and Isotropy
MRR 4.1-4.2
See introductory slides.
1
1 — Introduction:
Homogeneity and Isotropy
2 J. R. Pritchard
Chapter 2
2.1 Expansion
Before we start discussing a physical and mathematical model to describe an expanding
Universe, it is useful to be able to picture just what we mean by “expansion” in this
context. On “small” scales, corresponding to relatively nearby galaxies, it is completely
appropriate to think of the expansion as a relative velocity between any two galaxies,
increasing with time and distance. On larger scales, however, such that the time between
the emission of light at the distance galaxy and its reception here is large, then we
are really observing the galaxy at a very different time and we cannot say that we are
observing a quantity having anything to do with its velocity relative to us. Hence, we will
see that there are very many galaxies in the distance Universe for which we would naively
define a velocity v > c, but this is just not the correct way to interpret our observations.
A similar confusion occurs when talking about the distance to that galaxy. We sim-
ply cannot lay out relativistic meter sticks between us, now, and the distant galaxy,
then, which is the appropriate thought-experiment to do in order to define the distance.
Hence we will also find that there are a myriad of different ways to define the distance,
each numerically different, although none of them is any more correct than the others.
Nonetheless even at this early stage it is very useful to lay down a coordinate system in
our expanding Universe. We could of course imagine a proper coordinate system, with
the coordinates given by the number of actual, physical meters, from some fiducial point
which we could choose to be the origin. But in this coordinate system, due to expansion,
the galaxies are moving. Instead, try to picture a coordinate system with the grid lines
expanding along with all of the objects in the Universe. If we could freeze time, this
coordinate system would just look like a normal cartesian (or polar, or however we decide
to draw our axes), but the grid lines are further apart from one another in the future,
and closer together in the past. So for this case we don’t really want to measure things
in physical units (meters or megaparsecs), but really just numbers. Conventionally, how-
ever, we do choose the units so that they are equivalent to physical units today — since
many of our measurements are done now.
Real galaxies or other objects may have small movements (called “peculiar velocities”)
with respect to this coordinate system, but on average — due to the homogeneity and
3
2 — The Expanding Universe
isotropy of the Universe — they will be at rest with respect to these coordinates as the
Universe expands. We call this a comoving coordinate system, and hypothetical observers
expanding along with these coordinates are called comoving observers. We show this (in
two dimensions) in Figure 2.1.
where t0 is some arbitrary time, which we can take to be the present day. We can combine
these to get
rij (t0 ) = a−1 (t1 )rij (t1 ) = a−1 (t2 )rij (t2 ) = const . (2.2)
(It’s a constant since the value at any specific time is fixed.) So, at a general time, t,
or
ȧ ṙij
= . (2.5)
a rij
This equation looks trivial, but note that the right-hand side depends on i and j, and
hence which two points you’ve chosen, whereas the left-hand side doesn’t. Therefore, the
4 J. R. Pritchard
2.2.: The Scale Factor and Hubble’s law
3
r13(t2)
r13(t1) r23(t2)
r23(t1)
1 2
r12(t1) r12(t2)
1 2
time t = t1 time t = t2
scaling between any two points is just a function of time, and we can write most generally
ȧ ṙ
= (2.6)
a r
for an arbitrary distance r. (Note that we started our derivation assuming that a has no
units, but in fact in this last equation we see that we could also give a units of length
which is sometimes done.)
The quantity a(t) — the scale factor — is going to be with us throughout the
course.1 It describes the evolution of the Universe, and for a homogeneous and isotropic
Universe, it tells us almost everything we need to know. We’re also going to continually
encounter the combination ȧ/a, which is the expansion rate (also referred to as the
Hubble parameter ) and we will sometimes write
ȧ
H(t) = (2.7)
a t
and in particular
ȧ
H0 = H(t0 ) = (2.8)
a t=t0
is the expansion rate today, one of the most important quantities in cosmology. Today,
then, we can rewrite this equation as
These are all relative distances (recall the ij indices we’ve got rid of), but by convention
we can take one of the points to be the location of the earth from which we observe, and
the other to be the location of some galaxy (since we can only measure the distances to
objects, not arbitrary points!), so we can then write,
v = H0 d (2.10)
1
In some texts, including MRR, the scale factor is denoted R(t) rather than a(t).
Cosmology 5
2 — The Expanding Universe
where v is the velocity of the galaxy relative to us, and d is its distance. This is Hubble’s
law, which we’ve derived assuming homogeneity, but in fact Hubble found it through
observations as we discussed last time.
Comments about there not being a true centre. . .
v
ρ F
In addition to the usual F = ma, and Newton’s law of gravitation, we’ll need to
remember two very important corollaries from Newtonian gravity. First, there is no
gravitational field inside a spherical shell of matter. Second, we can treat the gravitational
attraction of a sphere as if it were concentrated at the centre. Hence, if we’re inside a
6 J. R. Pritchard
2.3.: Newtonian Dynamics
homogeneous sphere of matter of density ρ (assume it’s gas so we can move around)
at some distance r from the centre, we can ignore everything at greater radii. We will
also assume that the matter is pressureless so it takes no P dV work for it to move —
this is conventionally called “dust” or simply “non-relativistic matter” in cosmology and
relativity.
This is the setup for the Universe, as shown in Figure 2.3, except that we’ll take the
outer sphere to be arbitrarily (infinitely) large. This is a bit of a cheat in Newtonian
gravity, but in fact this derivation is correct, and indeed goes through to GR. So consider
a point P at a distance r from the “centre”. The gravitational force on a test particle of
mass m at P is
GM m Gm 4 4πGmρr
F = mr̈ = − 2
= − 2 πr3 ρ = − , (2.11)
r r 3 3
using the mass interior to r, M = 4πr3 ρ/3. We can rewrite this as
r̈ 4
= − πGρ (2.12)
r 3
Now, if we assume that the sphere is expanding homogeneously, our test mass at point
P is actually moving with time, r(t) ∝ a(t), but so is all the matter both interior and
exterior to it. Hence, although the density, ρ, of the interior sphere is changing with time,
the total mass in that sphere (M , above) remains constant. So in fact we can just go
back to the second equality above and write
MG
r̈ = − (2.13)
r2
where now we note that M is a constant with time. This is a differential equation that
we can integrate, and find
ṙ2 = 2GM/r + A (2.14)
where A is an integration constant. We can rewrite this in the form
2
ṙ 2GM A 8πG A
= − = ρ − (2.15)
r r3 r2 3 r2
where we have replaced M = 4πr3 ρ/3. With this substution, Eq. 2.12 and Eq. 2.14 only
contain factors of r in combinations ṙ/r, r̈/r or with an integration constant. Recall that
we can relate the distance between any two points as a function of time, r(t), to the scale
factor via r(t) = a(t)r(t0 ). Hence in these terms the factors of r(t0 ) cancel or can be
absorbed into a constant, and so we can rewrite these differential equations as
ä 4
= − πGρ (2.16)
a 3
and 2
ȧ 8πG kc2
= ρ(t) − 2 (2.17)
a 3 a
Cosmology 7
2 — The Expanding Universe
where we have written the integration constant as kc2 where c is the speed of light so that
the combination a2 /k has units [length]2 — some authors use that to make k a number
and give a units of length, while others make a a number and k have units [length]−2 .
For now only the overall combination k/a2 matters.
Exercise: solve Eq. 2.13 to get Eq. 2.14. This is a slightly simplified version of the
Friedmann Equation, and is one of the most important equations in cosmology. We
will encounter it again and again over the entire course. Despite the Newtonian derivation
(which required some hand-waving) it is in fact generally true for a homogeneous and
isotropic space-time, and was first derived by Friedmann using General Relativity.
We have used the fact that the mass in a sphere that expands along with the Uni-
verse remains constant. This is called a comoving sphere, another concept that we will
encounter again and again. We can use this observation to write down an equation for
the evolution of the density:
−3M dr
d d M ȧ
ρ(t) = = = −3ρ(t) = −3Hρ (2.18)
dt dt 4πr(t)3 /3 4πr4 /3 dt a
or
ρ̇ + 3Hρ = 0 (2.19)
a(t1 )3
ρ(t) ∝ 1/a(t)3 or ρ(t) = ρ(t1 ) (2.20)
a(t)3
where t1 is some arbitrary but fixed time. Plugging this into the Friedmann Equation,
2
ȧ 8πG a(t1 )3 kc2
= ρ(t1 ) − (2.21)
a 3 a(t)3 a(t)2
Up until now, we haven’t said much about a, but in fact we know that our Universe
is expanding (a is growing, or ȧ > 0), so if we go back far enough, a gets smaller and
smaller. Eventually, then, the first term above will dominate. (Alternately, we can look
at the case k = 0 which we will see is a very special case, and in fact may describe the
8 J. R. Pritchard
2.4.: Thermodynamics
where in the last line we have assumed a(0) = 0 (which we can take to define t = 0 —
when the distance between any two points was zero).
As an aside, there’s another interesting case, if k > 0 in Eq. 2.17 there’s a time when
ȧ = 0; the Universe stops expanding! We will see soon that this case corresponds to a
“closed” Universe which expands to a finite size and then recollapses.
2.4 Thermodynamics
MRR 4.3
If we go back to our conservation equation, Eq. 2.19, we can write it in an illuminating
form, multiplying the left and right sides by a3 dt:
a3 dρ + 3ρa2 da = 0
d(ρa3 ) = 0 ∝ dE/c2 (2.23)
where in the last line we have first used the fact that, for any comoving volume, ρa3 is
proportional to M , the mass in that volume, and then used relativity (E = M c2 ) to write
that mass in terms of the total energy in the volume. This is really just the first law
of thermodynamics — energy conservation — since for this “dust” matter we have both
dQ = T dS = 0 and dW = P dV = 0. But let’s generalize to P 6= 0:
dE = T dS − dW = 0 − p dV
dE/c2 = −p dV /c2 ∝ −p d(a3 )/c2 (2.24)
Cosmology 9
2 — The Expanding Universe
10 J. R. Pritchard
2.5.: General and Special Relativity
where the subscript 1 refers to some fixed time at which the quantities are evaluated. We
can solve this:
r
8πG 4
ȧ = a1 ρ1 a−1
r 3
8πG 4
a da = a1 ρ1 dt
r 3
2 8πG 4
a /2 = a ρ1 t (2.37)
3 1
or
a ∝ t1/2 radiation (2.38)
which we can compare to
where from now on we will usually refer to the “dust matter” as “non-relativistic matter”.
Cosmology 11
2 — The Expanding Universe
and
ä 4πG
ρ + 3p/c2 ,
=− (2.41)
a 3
along with the conservation equation
dρ p ȧ
+3 ρ+ 2 =0 (2.42)
dt c a
But the real Universe is governed by General Relativity (GR). So to really under-
stand what these equations mean (and in particular the significance of the constant of
integration, k) we need to use a little bit of GR.
where the indices µ, ν run over 0,1,2,3, corresponding to time and three spatial coordi-
nates, gµν is the metric, and dxµ is a differential spacetime coordinate interval (in order
to have units of length, we will take the time coordinate, µ = 0, to be dx0 = c dt where c
is the speed of light). Because we are free to use different coordinates (for example, polar
vs. Cartesian), the actual form of the metric depends on the coordinates used — and the
requirement that the physics not depend on this is one of the central facts of GR.
Note that we will occasionally use the “Einstein Summation Convention” and write
this as
ds2 = gµν dxµ dxν summed over repeated indices . (2.44)
You have probably already encountered a metric in special relativity, which we used
to describe the fact that the time coordinate can mix with the spatial coordinates. In this
case, the appropriate metric is the Minkowski metric, which, for Cartesian coordinates
(x0 , x1 , x2 , x3 ) = (ct, x, y, z) is just
+1
−1
gµν = ηµν = (2.45)
−1
−1
[note that we will use the “signature” (+ − −−) but you will often encounter (− + ++)
in the literature]. Here, then,
12 J. R. Pritchard
2.5.: General and Special Relativity
If instead we had decided to use polar coordinates, we could instead write the spatial
three-dimensional part as
where, finally dΩ22 represents the metric on the surface of the r = 1 sphere (more specifi-
cally, of the two-sphere, where “two” represents the number of dimensions of the surface
of the sphere) — this is our first example of the metric of a curved surface, even though
it is a surface embedded in a larger, flat, spacetime.
To get a feeling for this, let’s take a look at a metric that describes a circle — that
is, distances along the perimeter of a circle of radius R. This is just a one-dimensional,
curved manifold. It is pretty clear that we should be able to write
which simply says that distances along the perimeter of the circle satisfy ds = R dθ. But
it is also case that we can write the metric in a different coordinate system as
dy 2
ds2 = . (2.50)
1 − ky 2
This is clear if we make the substitution ky 2 = sin2 θ, in which case we can write
k −1/2 cos θ dθ
ds = p = k −1/2 dθ (2.51)
2
1 − sin θ
so we can identify k = 1/R2 and we can also interpret our second set of coordinates to
mean that y = R sin θ represents y in a Cartesian coordinate system (with θ measured
counter-clockwise from y = 0, x = R).
Exercise: The above discussion assumes k > 0. Show that k = 0 corresponds to
distances along a line, and k < 0 to distance along a hyperbola.
Cosmology 13
2 — The Expanding Universe
anywhere. It turns out there is one other, not so obvious, possibility, which is the “hy-
perbolic paraboloid” or saddle shape. All of these possibilities are shown in Figure 2.4.
For the sphere, we can obviously label the manifold by its radius in some units, and in
fact the flat plane corresponds to the infinite-radius limit. We can also define a “radius of
curvature” for the hyperbolic manifold. (Indeed, many calculations on the hyperboloid
just correspond to using the hyperbolic trigonometric functions in place of the spherical
functions.) For both of the curved cases, the manifold is self-similar: it is just a uniform
scaling between different values of the radius of curvature.
Figure 2.4: Homogeneous and isotropic manifolds in two dimensions. Left: All
three constant-curvature manifolds, courtesy WMAP/NASA GSFC. Right: Hyperbolic
paraboloid courtesy Wikipedia: http://en.wikipedia.org/wiki/File:Hyperbolic_
triangle.svg
14 J. R. Pritchard
2.5.: General and Special Relativity
follows geodesic paths, we can use lightrays, and therefore astronomical observations, to
probe the geometry of the Universe. Later on, we will use this very fact to measure the
curvature of the Universe as a whole with the Cosmic Microwave Background radiation.
It turns out that in 3 + 1 dimensions, we have the same set of homogeneous and
isotropic manifolds, but promoted to higher dimensionality (and so effectively impossible
to picture with our 3D brains). The most general spatially homogeneous and isotropic
manifolds are indeed flat space, the three-sphere, and the three-hyperboloid. Note in
particular that the three-sphere is not a “ball” in three dimensions — it is a three-
dimensional “surface” with no boundary that can be observed in three dimensions, just
as there is no boundary as you walk along the surface of the approximately-spherical
earth. The metric for this case is given by
dr2
2 2 2 2 2 2 2 2 2
ds = c dt − a (t) + r dθ + r sin θ dφ (2.52)
1 − kr2
where we use polar coordinates in three dimensions, and we now have a scale factor a(t),
since all that the cosmological principle tells us is that at any given time the Universe
is homogeneous and isotropic, but it lets us change the overall scale of the universe as a
function of time. In this form, the scale factor has no units. We also have a number k in
our equation, and this tells us which of the three possible manifolds we have. As we have
written it here, this is a real number with units [length]−2 , just as in the case of the circle
manifold. So in fact it is just the overall sign of k that determinesp which of the manifolds
applies. [You can p see this by making the substitution r → r/ |k| in which case we
can rescale a → a/ |k| as well, and only the overall sign of k matters. Because of this
freedom, you will sometimes see this metric written with the terms and factors scaled
in this way.] This is the famous Friedmann-LeMaı̈tre-Robertson-Walker metric (FLRW,
although sometimes LeMaı̈tre is left out to give only FRW). Amazingly, the scale factor
a and the constant k that we have been using in our Newtonian calculation correspond
to exactly the same quantities in this full analysis.
We can examine the spatial part of this metric separately:
dr2
dχ2 = + r2 dθ2 + r2 sin2 θ dφ2 . (2.53)
1 − kr2
We have removed the common a2 (t) factor which converts to physical coordinates; this
is the metric of a “comoving” constant-time spatial slice of the metric. This looks a bit
like some of the metrics we considered in the previous Subsection 2.5.1: it shares the
dr2 /(1 − kr2 ) term with the circle (exchanging r for y), and the remaining two r2 dΩ2
terms with the spherical manifolds. Combining these, for k > 0, gives the metric of a
three-sphere,
√ that is a three-dimensional manifold all of whose points are a distance
R = 1/ k from a fixed point in higher dimensions (I emphasize that this is not a normal
two-sphere). The discussion above tells us in particular that for k 6= 0 the r coordinate
does not represent distances along that sphere, but can be interpreted as a Cartesian
coordinate in the higher-dimensional space. The r coordinate does, however, represent
an important distance: it gives the radius (in comoving length) of a sphere centred at
Cosmology 15
2 — The Expanding Universe
r = 0. This is evident when we go back and recall the interpretation of the dr2 term
as the metric along the circle. In two dimensions we can think of a sphere as a series
of “nested” circles of radius r (i.e., as if looking down from the top at lines of constant
latitude), and in three dimensions of nested spheres of radius r. Note that the dr term
doesn’t give the metric along each of these spheres or circles, but how they are related to
one another.
The possibilities for the sign of k thus each correspond to one of our constant-curvature
manifolds. The choice k = 0 corresponds to the flat manifold, which should be clear from
the form of the metric, which in this case just looks like Minkowski with the extra factor
of a2 on the spatial part. So at any one time, the Universe acts like Minkowski space,
although of course anything that actually happens takes a finite amount of time, so the
evolution of the Universe must be taken into account.
If k > 0, we say that the Universe has positive curvature, and it takes the form of
a three-sphere at any given time. Analogously to the two-dimensional sphere discussed
above, triangles have a total of more than 180 degrees, lines that start out parallel will
converge and eventually cross, any seemingly-straight line returns to its starting point
eventually, and the Universe has a finite volume at any one time. This corresponds to the
case discussed a few lectures ago where we saw that a0 (t) = 0 at some time, corresponding
to a maximum expansion followed by recollapse: the sphere grows with time from R = 0
and then shrinks.
If k < 0, we say that the Universe has negative curvature, and it is the three-
dimensional saddle shape or hyperboloid. Here, a triangle has less than 180 degrees,
and lines that start out parallel may cross and will diverge, but, like the flat case, the
Universe is infinitely large.
16 J. R. Pritchard
2.5.: General and Special Relativity
the left-hand side, where gµν is the metric, and Λ is the infamous cosmological constant
(“Einstein’s biggest blunder”) — and it really is a constant, a single number given once
and for all. In fact, we’ll see that even this term is surplus to requirements and in fact
can really be subsumed into the stress-energy tensor itself, and despite being a “blunder”
is probably needed to describe the Universe. So, if we include this term, the Friedmann
Equations become 2
ȧ 8πG k 1
= ρ(t) − 2 + Λ , (2.55)
a 3 a 3
and
ä 4πG 1
=− ρ + 3p/c2 + Λ . (2.56)
a 3 3
In these equations, ρ, p and a are all functions of time, and they must be supplemented
with enough information about the behaviour of the mass-energy to solve these equations
(for example, fluid conservation, Eq. 2.42 along with the equation of state parameter,
wc2 = p/ρ).
2.5.3 Redshift
MRR 3.3, 7.4
Almost all of the information we have about the Universe comes to us in the form
of light — photons — that have propagated to us from distant objects. Because of the
finite speed of light, of course, this means that we see objects as they were at some time
in the past (and it also means that defining the distance to an object that you see can
be complicated). Light travels along the geodesic curves of a manifold, essentially locally
straight lines, satisfying ds = 0.
So let’s return to our FRW metric, Eq. 2.52, and look at the case where light travels
along a “radial” curve with constant θ and φ (so dθ = dφ = 0), from the point of emission
at t = te , r = re to our observation today at t = t0 , r = r0 = 0. We then have
dr2
0 = ds2 = c2 dt2 − a2 (t) (2.57)
1 − kr2
or
c dt dr
= −√ (2.58)
a(t) 1 − kr2
where we have chosen the negative sign since both a and t increase to the future, whereas
r increases away from the observer, i.e., toward the past. We can integrate this along the
path from emission to observation:
Z t0 Z 0 Z re
c dt −dr dr
= √ = √ = fk (re ) (2.59)
te a re 1 − kr 2
0 1 − kr2
where we note that the dr integral is a fixed function of re (and k). Because of this,
the equality must still hold for photons emitted at a later time, te + δte and observed at
t0 + δt0 (note that the two δs are different). The δs correspond to, for example, successive
Cosmology 17
2 — The Expanding Universe
peaks in the lightwave, since we’ll see that we want to measure the photon wavelength
or frequency. Hence Z t0 Z t0 +δt0
dt dt
= (2.60)
te a te +δte a
We can now split up the range of integration on both sides to eliminate the “overlap”
between t0 + δt0 to te :
Z te +δte Z t0 Z t0 Z t0 +δt0
dt dt dt dt
+ = + . (2.61)
te a te +δte a te +δte a t0 a
The two “inner” terms are equal, so we are just left with
Z te +δte Z t0 +δt0
dt dt
= . (2.62)
te a t0 a
Now, let’s assume that the δs are small and Taylor expand the integrands:
1 1 ȧ 2 1
= 1 − δt + O(δt ) ' (2.63)
a(t + δt) a(t) a a(t)
where the approximation holds if
δt c−1 × photon wavelength
' 1 (2.64)
(ȧ/a)−1 age of Universe
and we’ve used H = ȧ/a ∼ 1/t0 and the fact that we’re considering successive peaks in
the sinusoidal lightwave. This inequality obviously holds for any reasonable times. Now,
applying this to t = t0 and t = te ,
δt0 δte
= (2.65)
a(t0 ) a(te )
which we can rewrite as
a(t0 ) δt0 λ0 /c νe
= = = (2.66)
a(te ) δte λe /c ν0
or
a0
=1+z (2.67)
a
where we define the redshift
νe νe − ν0 δν
z = −1= =
ν0 ν0 ν
λ0 − λe δλ
= = (2.68)
λe λ
What does redshift mean? Most importantly, it means that the wavelength of freely-
propagating photons increases with the expansion of the universe, proportional to the
scale factor. This is a purely General-relativistic effect. Recalling our heuristic Newtonian
18 J. R. Pritchard
2.5.: General and Special Relativity
derivation of the expansion of the Universe, we can see that the expansion is very much
like climbing out of a potential well, and so it is plausible that moving objects — photons,
in this case — would lose energy as they propagate. Note that although our derivation
required approximations when we used the Taylor expansion, in fact the cosmological
redshift of Eq. 2.67 is exact in a full GR calculation.
We also see that when z > 1 (which we certainly do observe for sufficiently distant
objects), if we use the usual correspondence between redshift and speed, v = cz, we
seem to have a relative speed greater than that of light. On the one hand, there is no
problem here: the equations that give this result are unambiguous. But what does it
mean? We are taught that one of the underlying principles of relativity is v ≤ c, and
yet we seem to contradict this. The answer is that this is a matter of interpretation,
not physics: Is v = cz correct in this circumstance? What does the prohibition against
superluminal velocities mean for very distant objects? We will not go into the details here,
but I will point to a recent paper by two of my colleagues, Bunn and Hogg, available
at http://arxiv.org/abs/0808.1081 and discussed further in Bunn’s blog at http:
//blog.richmond.edu/physicsbunn/2009/12/02/interpreting-the-redshift/.
Interpretation aside, the redshifting of radiation as the Universe expands has some im-
portant and interesting effects. Let us compare photons to our “dust” or “non-relativistic
matter”. As the Universe expands, the number density of matter particles decreases:
nm = N/V ∝ N/L3 ∝ a−3 . Hence the energy density, which is just ρm c2 = mnm c2 for
particle of mass m, also scales as
ρm c2 ∝ a−3 . (2.69)
For photons, the number density is still n ∝ a−3 . However, the energy per particle
is Eγ = pc = hν = hc/λ ∝ a−1 as the wavelength redshifts. Hence, for photons,
ργ c2 = nγ Eγ ∝ a−3 × a−1 or
ργ c2 ∝ a−4 . (2.70)
So far, we have discussed the case of photons, but really this distinction is more general.
Consider the fully relativistic energy per particle, E 2 = p2 c2 + m2 c4 . If the first term
dominates, (mc2 pc), the particles behave like photons, which we will call, generically,
radiation — this applies to anything with zero mass, but also to very low-mass particles,
where “low-mass” really means “high-speed”, since in the early Universe we will see that
all particles move with greater and greater speed as the Universe gets hotter and hotter
as we look earlier and earlier. Conversely, when the second term dominates (mc2 pc),
their energy is dominated by their mass, and they behave like non-relativistic matter.
We will very often use the redshift as a proxy for the time parameter — if the Universe
has always been expanding, then the redshift of a more distant object (hence with an
earlier te is always greater). Obviously, this is the redshift as observed by us, today; it
would have a different value if observed at a very different time. (As we can see from our
derivation, however, as long as the time difference is small compared to the age of the
Universe — which it will be for any observations made in the recent past or foreseeable
future! — the redshift will be the same. We may explore this further in a problem set.)
Cosmology 19
2 — The Expanding Universe
20 J. R. Pritchard
Chapter 3
Consider what this equation looks like if we evaluate everything today, t = t0 . We define
the Hubble Constant as
ȧ
H0 = = 100h km s−1 Mpc−1 (3.2)
a t=t0
where the second equality just parameterizes our ignorance: we are pretty sure from
observations that H0 is between 50 and 100 km s−1 Mpc−1 , so this way the quantity
that we don’t know is 0.5 <
∼h< ∼ 1, which is a bit easier to work with. With this, the
Friedmann equation is
8πGρ0 kc2 1
H02 = − 2 + Λ. (3.3)
3 a0 3
Now, consider a flat Universe with no cosmological constant, k = Λ = 0, in which case
we will say that the density today is the “critical density”, ρ = ρc , given by
8πGρc
H02 = (3.4)
3
or
3H02
ρc = = 1.9h2 × 10−29 g cm−3 . (3.5)
8πG
21
3 — Cosmological Models and Parameters
Even for a universe with ρ 6= ρc , we can define the ratio of the actual density to the
critical density, the density parameter
ρ0
Ω0 = (3.6)
ρc
The two parameters H0 and Ω0 have traditionally been the most important numbers in
modern cosmology, and until the last decade or so, their values were not really known
to better than 50% or so. However, we’ll see right away that we can’t really use only a
single number, Ω0 , to represent the density; we really a need a separate number for each
“kind” of matter.
To see this, let’s rewrite the Friedmann equation in a very useful form. First, let’s
write the expansion rate as
ȧ
H(t) = = H0 E(z) (3.7)
a
which defines a new function E(z), first popularized by James Peebles in his books and
papers, which we will call the “Hubble Function”. The Friedmann equation becomes
8πGρ kc2 Λ
E 2 (z) = 2
− 2 2
+
3H0 H0 a 3H02
ρ(z) kc2 Λ
= − 2 2 (1 + z)2 + . (3.8)
ρc a0 H0 3H02
From the previous chapter, we know that, for non-relativistic matter,
−3
a
ρm (z) = ρm (0) = ρm (0)(1 + z)3 = ρc Ωm (1 + z)3 (3.9)
a0
and for radiation
−4
a
ρr (z) = ρr (0) = ρr (0)(1 + z)4 = ρc Ωr (1 + z)4 . (3.10)
a0
We can similarly define
kc2 Λ
Ωk = − and ΩΛ = (3.11)
a20 H02 3H02
(we really shouldn’t think of Ωk , especially, as having anything to do with an energy
density, although we will see that ΩΛ actually is related to an appropriate energy density
ρΛ c2 ), which finally gives
E 2 (z) = Ωm (1 + z)3 + Ωr (1 + z)4 + Ωk (1 + z)2 + ΩΛ . (3.12)
In particular, at z = 0 (today), E(0) = 1, since ȧ/a|0 = H0 , so
1 = Ω m + Ωr + Ωk + ΩΛ (3.13)
where I will always use the quantities Ωi to refer to the value today. Note that this
assumes that Eqs. 3.9-3.10 are exact — that there is no conversion between components
due to particle decay, for example. Because all of the Ωi except Ωk can refer to the actual
density of some sort of “stuff”, we shall also define the total density parameter
Ωtot = Ωm + Ωr + ΩΛ = 1 − Ωk . (3.14)
22 J. R. Pritchard
3.2.: Cosmological Models
Because the evolution at this time is controlled by the value of Ωm , we say that the
Universe is matter-dominated (MD). However, if we go to a somewhat earlier time (any
z 0.3/2.5 × 105 ), the Ωr term is largest and we say that the Universe is radiation-
dominated (RD). Similarly, if we wait long enough, and if Ωk and ΩΛ are nonzero, we
expect them to dominate eventually. We show all of these in Figure 3.1, although of course
the real universe wouldn’t have such sharp transitions between the different phases. Note
that the actual Universe need not have all of these phases, if one of the Ωi is sufficiently
small or zero. In fact, we’re pretty sure that Ωk = 0 in our Universe, so the curvature-
dominated phase probably doesn’t happen. Moreover, it looks like ΩΛ ' 0.7, so we are
experiencing the change-over from MD to ΛD now.
Cosmology 23
3 — Cosmological Models and Parameters
-4
RD ρ∝a
log ρ
-3
MD ρ∝a
-2
CD ρ∝a
0
ΛD ρ∝a
aEQ
log a
Figure 3.1: The evolution of the density of various components. RD, MD, CD, and ΛD
refer to Radiation-, Matter-, Curvature- and Λ-domination, respectively.
so
ȧ
= t−1 . (3.19)
a
This is the “wide open”, or Milne Universe. Note that it also corresponds to the far
future evolution of any open Universe (k = −1 or Ωk > 0 or Ωtot < 1) in which the
matter and radiation densities have diluted so much as to become negligible, as long as
there is no cosmological constant ΩΛ = 0.
a ∝ t2/3 (3.20)
which gives
ȧ 2
= t−1 . (3.21)
a 3
This corresponds to early times for a Universe with non-relativistic matter (but not so
early that a radiation component dominates), and is often called the “Einstein-de Sitter”
Universe (not to be confused with the “de Sitter” Universe, which is, unfortunately, a
different case!).
24 J. R. Pritchard
3.2.: Cosmological Models
The Closed Universe. Now, consider k > 0, Ωtot > 1, or Ωk < 0. We saw earlier
that in this case there is a value of a such that ȧ = 0. This occurs when
8πGρ kc2
= 2 (3.22)
3 a
or (assuming non-relativistic matter so ρ ∝ a−3 )
8πG
amax = ρ0 a30 . (3.23)
3kc2
In fact, the full solution for the evolution of a(t) is symmetric around this point: it climbs
from a = 0 at t = 0 to amax at t = tmax and then falls back to a = 0 at = 2tmax .
So far, we’ve ignored the possibility of a radiation-dominated Universe. In fact, in
terms of the long-term evolution of the Universe, it appears that the radiation-dominated
phase is very short and we are now long after it. Exercise: show that the effect on the age
of the Universe (i.e., the relationship between a and t) of an early RD phase is negligible.
However, we will see that much of the most important “microphysics” of cosmology
happens during the RD phase, so it is worth recalling its properties.
Cosmology 25
3 — Cosmological Models and Parameters
Figure 3.2: Left: The scale factor as a function of time, for MD universes with k < 0,
k = 0, k > 0. [Courtesy Ned Wright, http://www.astro.ucla.edu/~wright/cosmolog.
htm.] Right, the scale factor for various universes, marked by their values of Ωm (with
ΩΛ = 0 except as marked), normalized to have the same scale factor and expansion rate
today.
this is a constant — the expansion rate does not evolve with time. This has solution
hp i
a ∝ exp Λ/3 t . (3.27)
Unlike the cases we have looked at so far, which had power-law evolution for a(t), this
Universe grows exponentially. We will see that this has important consequences.
However, note that the early time behaviour of the Universe is almost completely
independent of the cosmological constant: the very early Universe looks like a flat RD
Universe, followed by an MD phase.
The Concordance Universe As noted, the real Universe seems to have matter,
radiation and a cosmological constant. Roughly,
Ωk = 0
Ωm ' 0.3
Ωr ' 10−5
ΩΛ = 1 − (Ωm + Ωr ) ' 0.7 . (3.28)
This is occasionally called the “standard cosmological model” or the “concordance cos-
mology”. More details on the current best measurements of these numbers (and others
you will encounter in this course) are at http://lambda.gsfc.nasa.gov.
26 J. R. Pritchard
3.2.: Cosmological Models
Note that in such a multicomponent Universe the behaviour can be quite complicated.
In Figure 3.2 we see this “ΛCDM Universe”; note that a(t) starts out decelerating but
(around now, t ' t0 ) the Cosmological Constant is beginning to dominate the expansion,
getting closer and closer to the exponential expansion of Eq. 3.27. Note also that this
enables the Universe to be considerably older than other possibilities, even an open
Universe with the same matter density today.
Usually, when we write Ωi we refer to the value today. But obviously this is just
a convenience to refer to quantities that we can measure easily. The density itself is
obviously a function of time, and we can define the critical density as a function of time,
as well, and so make the Ωi functions of time or redshift as well:
2
3 ȧ 3H02 2
ρc (z) = = E (z) = ρc,0 E 2 (z) , (3.29)
8πG a 8πG
so we can define
ρi (z) ρc,0 Ωi,0 (1 + z)ni
Ωi (z) = =
ρc (z) ρc,0 E 2 (z)
(1 + z)ni
= Ωi,0 (3.30)
E 2 (z)
where we’ve used a zero subscript to make sure it is evident we are referring to today,
t = t0 (but probably will not in the future). For a Universe with radiation, matter and
a cosmological constant, this is shown in Figure 3.3. In a flat Universe, the components
are Ωi = 0 or Ωi = 1 for much of the evolution of the Universe. This is not the case for
an open or closed Universe.
In fact the concordance Universe is a fairly “boring” example of a Universe with a
cosmological constant. As you will see in the problem sheet, if we balance the densities
of the various components just right, we can get a wide variety of behaviours (including
the one that led Einstein to his “greatest blunder”.)
Cosmology 27
3 — Cosmological Models and Parameters
W
1.0
0.8
0.6
0.4
Wm HaL
Wr HaL
0.2
WL HaL
Figure 3.3: The density dependence of the various components of the concordance Uni-
verse. Note that the horizontal axis gives the scale factor of the Universe, with a/a0 = 1
today.
28 J. R. Pritchard
Chapter 4
Cosmography
In this chapter we will discuss the description — and measurement — of times and
distances in a Friedmann-Robertson-Walker univese. How old is the Universe? How far
away is some object that we see? We will see that these questions do not necessarily
have unambiguous answers, and that we must very carefully describe what we want to
measure.
There are many ways to actually do this integral, but one particularly simple one is to
use the fact that
a 1
= so da = −a0 (1 + z)−2 dz (4.2)
a0 1+z
and we have defined
ȧ
= H0 E(z) (4.3)
a
so
Z ∞
dz
t0 = 2
(1 + z) [H0 E(z)]−1
0 (1 + z)
Z ∞
−1 dz
= H0 , (4.4)
0 (1 + z)E(z)
where E(z) is the Hubble Function of Eq. 3.12. (Since E(z) depends on the quantity
1 + z it may be useful to use 1 + z or 1/(1 + z) as an integration variable.) Exercise: Show
29
4 — Cosmography
that the time ∆t(z1 , z2 ) between two redshifts z1 and z2 is given by the same expression
with the limits of integration changed to z1 and z2 .
Remember that we saw (in a problem sheet) that if we had a constant expansion
velocity ȧ we would have found an age 1/H0 , so we can think of the combination
Z ∞
dz
H0 t0 = (4.5)
0 (1 + z)E(z)
as being entirely due to the change in expansion velocity, which is of course due to the
presence of matter in the Universe. For most of the models we will consider, the quantity
H0 t0 is of order one. For example, in a flat, matter-dominated Universe, H0 t0 = 2/3
Considering Figure 3.2, this is the difference between a tangent line stretching back
from today to the horizontal axis compared to the actual curve. In this figure, a(t) always
has a negative second derivative (expansion decelerates). This is because we only showed
Universes with no cosmological constant, but in fact once we allow ΩΛ > 0, we can have
ä > 0 which can actually make the Universe “older” than it would be with constant ȧ,
as we saw in the previous chapter, Figure 3.2.
This fact has been crucial in our understanding of our actual Universe. We find
from measurements that H0−1 ' 14 Gyr and in fact find objects in the Universe that we
believe to be roughly 11–12 Gyr old. Since this gives H0 t0 ' 0.8 >∼ 2/3, it seems that the
simplest FRW Universe — flat and matter-dominated — cannot be the case! Either our
measurements are wrong, or we need the Universe to be dominated by a form of matter
which gives a larger value of H0 t0 . But we have seen that this will have the effect of an
accelerating scale factor, or a positive value of ä. This seems a very strange thing when
we recall the equation that governs the acceleration of the expansion:
ä 4πG Λ
=− ρ + 3p/c2 + (4.6)
a 3 3
Obviously Λ > 0 will do this, but it is worth considering the strange-seeming possibility
that Λ = 0 but that the first term is overall positive — which means that somehow
ρ + 3p/c2 < 0.
4.2 Horizons
MRR 4.10
Information cannot travel faster than the speed of light, or more precisely causality
cannot act faster than the speed of light. This is one of the core principles of physics
since the early twentieth century, affirmed in relativity and quantum mechanics alike.
Light rays travel on null geodesics, which in our metric notation corresponds to ds2 =
0. If we assume dφ = dθ = 0 as in our redshift calculation of Section 2.5.3. We can first
ask, at what distance would a light ray have had to start to arrive here at time t? We
show a spacetime diagram of this situation in Figure 4.1.
30 J. R. Pritchard
4.2.: Horizons
ae × χe
Past A B a=a(te)
"Light
Cone"
Expansion
Present a=a0
A B
a0 × χe
where dχ2 gives the metric of a three-dimensional spatial “slice” of spacetime, in comoving
coordinates, and dΩ2 = dθ2 + sin2 θdφ2 gives the angular metric. We need to then use
the same equation as for the redshift calculation, for a light ray with dθ = dφ = 0:
c dt dr
= −√ . (4.8)
a(t) 1 − kr2
Note that this equation is just dχ, a differential element of comoving distance. As in the
redhift case, we put t = 0 at r = rH , and t = t0 at r = 0 and hence a negative sign
appears. We can integrate this along the line of sight from t = 0, r = 0 to t, r = rH , after
switching the limits of integration because of that negative sign (equivalently, we could
just use r = 0 at t = 0 and r = rH at t):
Z t Z rH
c dt dr
= √ ≡ χH (t, rH ) (4.9)
0 a(t) 0 1 − kr2
where we define χH , the comoving distance to the horizon at comoving coordinate value
rH . For a flat Universe (k = 0) these two are the same. However, for k 6= 0, they are
Cosmology 31
4 — Cosmography
not, and in particular, χH corresponds to the comoving distance to the sphere — along
a t = const surface — defined by coordinate rH . As we discussed in Section 2.5, r does
not correspond to a distance along the manifold; for example in the spherical case it is
basically the distance perpendicular to the axis defining r = 0, or alternately the radius
of the two-sphere labelled by r. (To see this, think of the r coordinate as referring to the
distance from the axis to a constant latitude circle of a 2-sphere.)
Finally, we convert this coordinate distance to a proper distance to the horizon by
multiplying by the scale factor, a(t),
Z t
c dt0
dH (t) = a(t)χH (t) = a(t) 0
. (4.10)
0 a(t )
We can recast this in a useful form by returning to our Hubble function, E(z), defined
by ȧ/a = H0 E(z), so
Z a Z a 0 Z ∞
dt a0 0 a 1 0 −1 a dz 0
dH (t) = ca 0 02
da = ca 0 02
da = cH0
0 da a 0 ȧ a a0 z E(z 0 )
∞
cH0−1 dz 0
Z
= . (4.11)
1 + z z E(z 0 )
In particular, the horizon distance today R is just this3/2evaluated at z = 0. For a flat,
−1 ∞ −1
Ωm = 1 Universe, we have dH = cH0 0 dz/(1 + z) = 2cH0 = 3ct0 where we have
done the integral and then used H0 t0 = 2/3 for this case from above. Note that this is
greater than the naive expectation of dH ∼ ct0 , the simple distance traveled by a photon
in the time from t = 0 to t = t0 , as dH is the distance today from r = 0 to the present-day
position defined by coordinate r = rH . As we see in the Figure 4.1, expansion means
that this position is further away than ct0 .
4.3 Distances
It is often useful to be able to describe the distance between different points (events)
on a spacetime manifold, but in GR there is usually no unambiguous definition of the
distance between two points. Rather, we must state very explicitly what we mean. If
we could freeze the Universe at a particular time, it might make sense to use the spatial
coordinate distance between those two points, and indeed that is an easy quantity to
work with. However, in practice we can only observe points that are on our “light cone”
— at a given redshift we observe objects at a particular coordinate distance from us at
a particular time in the evolution of the Universe. We have already seen this in our
discussion of the redshift. Since a geodesic (light ray) satisfies ds ≡ 0, we have
dr2
0 = ds2 = c2 dt2 − a2 (t) (4.12)
1 − kr2
which we can reorder and integrate along the path from emission to observation (again,
we can either define r backward along a light-ray coming toward us or vice versa):
Z t0 Z re
c dt dr
χe ≡ = √ = fk (re ) (4.13)
te a 0 1 − kr2
32 J. R. Pritchard
4.3.: Distances
which just converts to distance along the circle, χe , from the “Cartesian coordinate”
which r represents as we saw when discussing the geometry of the circle in Section 2.5.1.
Recall that we had k 2 = 1/R where R is the radius of curvature (which is the actual
radius of the sphere for k > 0) so these make dimensional sense.
We can solve this equation for re ,
p
−1/2
|k| sin |k| χe k>0
re = χe
p k=0
|k|−1/2 sinh
|k| χe k<0
p
≡ |k|−1/2 sink |k| χe
c −1/2
p
−1
= |Ωk | sink |Ωk |a0 H0 c χe , (4.15)
a0 H0
which we writep in the unified form defining sink for simplicity (and taking sink (x) = x and
cancelling the |k| factors when k = 0 — we can also think of this as the k → 0 limits
of the positive and negative k cases). We have used the definition of Ωk ≡ −kc2 /(a0 H0 )2
to eliminate k. We can also put the χe integral into a form we can use more simply:
Z t0 Z t0 Z ze Z ze
−1 dt dt da dz a 1 dz
a0 H0 c χe = a0 H0 = a0 H0 = a0 H0 = (4.16)
te a te da a 0 1 + z ȧ a 0 E(z)
[Note: very often, cosmologists will absorb R into the definition of the r coordinate or
the scale factor a and take k = ±1 or k = 0. Dimensionally, the form of the FRW metric
dictates that kr2 has no units (from the 1 − kr2 factor) and that ar has units of length.
This can be realized by giving a units of length and making both k and r unitless, or
by making a unitless, giving r units of length, and k units of length−2 . There is further
freedom of scaling (i.e., units): very often you will see the convention a0 = 1 so that
distances are measured as compared to the present day. Note that finally we write these
equations in terms of Ωk , which is unambiguously defined.]
Cosmology 33
4 — Cosmography
by F = L/(4πd2 ) for an object at distance d. We can use this to define the “luminosity
distance”, dL , in a curved and expanding spacetime:
L
F= (4.17)
4πd2L
where dL can only depend upon the redshift once we have fixed the spacetime. To
calculate it, consider a spherical coordinate system with the origin at the source, S, at
radial coordinate re (so that our observation point is on the sphere), emitting at time te ,
observed at t0 . The total area of the sphere that goes through our observation point at
t0 is
Z Z
2 2 2 2
A = a0 re dΩ = a0 re d cos θ dφ = 4πa20 re2 . (4.18)
This equation uses the scale factor today, a0 , because this gives the proper, physical,
area of the sphere at t0 . Because of the expansion of the Universe and the redshift,
the total rate of energy coming through the sphere is modified from its rate at emission
by two effects. First, the number of photons coming through the sphere per unit time
is decreased by a0 /ae = 1 + ze . Second, each of those photons is doppler shifted so its
energy is decreased by an additional factor of a0 /ae = 1+ze . Note that these are different
effects, and hence the energy per unit time is decreased by (1 + z)2 . (Another way to
think of this is that the total energy, integrated over all time, would be decreased by one
factor of 1 + z due to redshift, but that energy is spread over a time that is longer by
another factor of 1 + z.) Hence,
L
F= (4.19)
4πa20 re2 (1 + z)2
so that
dL = a0 re (1 + z) . (4.20)
Now, we can combine this with our solution from above for re (Eqs. 4.15-4.16) to give
Z ze
p dz
dL (ze ) = cH0−1 (1 + ze )|Ωk |−1/2
sink |Ωk | . (4.21)
0 E(z)
These expressions do not depend on the characteristics of the source (L) nor on the
measured flux, but do depend on the cosmological parameters through the factors of H0
and E(z) in the integral, as well as k itself.
34 J. R. Pritchard
4.3.: Distances
is independent of the cosmology (for fixed physical size D and luminosity L).
In Figure 4.2 we show the behaviour of the luminosity and angular-diameter distances
as a function of redshift. Note that the angular diameter distance actually turns over:
this means that objects actually can start getting bigger as they get further away! Un-
fortunately, due to the dimming of the surface brightness (Eq. 4.26), they become harder
and harder to see, very rapidly, as the energy is spread out over a wider and wider area
of sky.
Cosmology 35
4 — Cosmography
Figure 4.2: Distance measures and the age of the Universe in different cosmologies.
Distances are measured in units of c/H0 and times in units of 1/H0 .
36 J. R. Pritchard
Chapter 5
MRR 5.1–5.2
Today, the evolution of the Universe is transitioning from being dominated by non-
relativistic matter to dark energy, while the temperature of the Cosmic Microwave Back-
ground is T ' 2.73 K (there are more energetic photons produced by astrophysical
processes as well). Because of this cold temperature, particle energies are low and in-
teractions are rare (except in places like stars, or CERN, of course). As we trace the
evolution of the Universe backwards, however, it heats up and the mean particle energy
increases, so interactions become more and more likely. Hence, we need to understand
the relationship between thermodynamics and particle physics in the early Universe,
sometimes called the “primeval fireball”.
or
37
5 — Thermodynamics and Particle Physics
(This number is probably uncertain to 10-20%.) At redshifts higher than zeq , the Universe
was radiation dominated. At this time, we have Teq = (1+zeq )T0 ∼ 104 K or kTeq ∼ 1 eV.
FIG. 4.ÈUniform spectrum and Ðt to Planck blackbody (T ). Uncertainties are a small fraction of the line thickness.
Figure 5.1: The measured spectrum of the Cosmic Microwave Background, plotted along
with a black body curve with T = 2.728 K. Uncertainties are a small fraction of the line
of the spectra derivedFrom
thickness. from the DIRBE
Fixsen templates
et al, can be
in the J., 473,
Astrophys. 576 (1996).on the unknown parameters p, G0 , and
performed
principal Ðt, g (l), and Ðt for their intensities at each pixel. *T . The Ðrst two terms are the Planck blackbody spectrum,
k
The CMBR temperature assumes a Planck spectrum. These with the temperature T ] *T . It is important to have the
spectra were chosen to match the shape of the spectrum in second term in order to0 properly estimate the uncertainty
The blackbody distribution of
the data. This yields maps of the CMBR temperature and
photons of frequency ν is given by the Planck function:
since the *T is strongly correlated with the resulting p (95%
dust intensities. A monopole plus three dipole components in the case of the Bose-Einstein distortion). The third term
are then Ðtted to the resulting temperature map. 8π allows
ν 2 dνfor Galactic contamination to remain in the mono-
The vector sum of the dipole coefficientsn(ν) points
dν in=the3 pole spectrum. The . Ðnal term is the modeled deviation.
(5.3) We
direction (l, b) \ (264¡.14 ^ 0.15, 48¡.26 ^ 0.15), consistent c exp(hν/kT ) −Kompaneets
Ðt either the 1 parameter or the chemical poten-
with the direction from the DMR results. Data for tial, but the two are too similar to Ðt simultaneously. The
o b o \ 10¡ Here,
were excluded from theconstant
h is Planck’s dipole Ðt because
so hν is of the uncertainties
the photon energy,are propagated
and the energyfrom distribution
the template Ðts, is and the
potential inaccuracy of the model of the Galaxy. The direc- correlation between the g(l) and LS /Lp increases the uncer-
therefore given by ε(ν) dν = (hν)n(ν) dν. c
tion is particularly sensitive to the Galaxy because it is tainty of G and p.
0
almost orthogonal to the direction of the Galactic center.
Galactic variations
38 in spectral shape as a function of longi- 6.1. Galactic Contamination
J. R. Pritchard
tude couple into the angle directly, while for the Ðxed angle Most of the Galactic emission has been removed, but
case, they come in as second-order terms for the same there is a small residual contamination. We use either the
reason. ; g (l) derived from the all-sky data set (see ° 4) or the
k (T k ), and we Ðt a temperature and an emissivity. The
l2B
l
l2B (T ) model with a temperature of 9 K produces a lower
5.2.: Black Body radiation and equilibrium statistics
Cosmology 39
5 — Thermodynamics and Particle Physics
where g is the number of degrees of freedom for the particle species (e.g., g = 2 for
spin-1/2 fermions). We can similarly calculate the energy density
Z
ε = g (p)n(p) d3 p (5.9)
as well as quantities like the average pressure (and hence the equation of state relating
density, temperature and pressure).
Note that if mc2 , the energy is dominated by the momentum term, just like for
photons. However, the majority of particles in either the BE or FD distribution have
energy ∼ kT . Hence, if kT mc2 , the mass of the particles are irrelevant — they
behave like radiation, and this is true for both bosons and fermions. (Strictly speaking,
we are assuming the chemical potential is much less than kT as well.)
In this relativistic limit, when we also assume µ m (non-degenerate), we find
(
π2 4 7/8 Fermions
ε=g (kT ) ×
30(h̄c)3 1 Bosons
(
ζ(3) 3 3/4 Fermions
n=g 2 (kT ) ×
π (h̄c)3 1 Bosons
ε
p= (5.10)
3
where the last line gives the pressure, and we use h̄ = h/(2π). The ratio of the energy
density to the number density gives the average energy per particle,
(
3.15 kT Fermions
hEi = ε/n ' (5.11)
2.70 kT Bosons
which we will very often just take to be hEi ∼ 3 kT .
This has important cosmological consequences. At earlier times, the temperature was
higher by the usual redshift factor T = (1 + z)T0 . For any particle of mass m, there is a
redshift such that kT > mc2 , before which the particle essentially behaves like radiation.
Hence, as we go backwards in time, the number density — and hence energy density
— of particles behaving like radiation (for example, having equation of state p = ε/3)
increases. Because the expansion of the Universe depends on the density of radiation-like
and non-relativistic-matter-like particles separately, we need to take this into account
when describing the expansion of the Universe at early times. The total energy density
requires summing Eq. 5.10 for ε over all relativistic particle species i, which as we will
see may each have a different temperature Ti , giving
π2
εr = g∗ (T ) (kT )4 (5.12)
30(h̄c)3
where we define the effective degrees of freedom as
4 4
X Ti 7 X Ti
g∗ = gi + gi (5.13)
i∈bosons
T 8 i∈fermions T
40 J. R. Pritchard
5.2.: Black Body radiation and equilibrium statistics
3/2
mkT 2 −µ)/kT
nP = g e−(mc (5.14)
2πh̄2
This is the number density in a proper volume element. If we want to know the number
density in a comoving volume element, we must rescale by the ratio of the comoving to
proper volume, proportional to a3 ∝ (1 + z)−3 ∝ T −3 .
In Figure 5.2 we combine these and show the comoving density as a function of
temperature (and hence redshift, since T = (1 + z)T0 ) for such a particle, assuming
µ = 0 (no degeneracy). For mc2 kT , this is a constant — the particle behaves
like radiation. For mc2 kT the proper density decreases exponentially as Eq. 5.14.
(Numerical integration of the Fermi-Dirac or Bose-Einstein distributions is required to
interpolate between the two limits.)
n¥
0.1
0.001
10-5
10-7
10-9
m c2
0.01 0.1 1 10 100 kT
Figure 5.2: The equilibrium comoving number density of a non-degenerate particle species
as a function of temperature. We plot the density relative to the number density at
T → ∞ (more properly, the m → 0 limit of Eq. 5.8) to remove the overall dependence
upon the particle properties and temperature. Both BE and FD densities are shown, but
they are nearly indistinguishable.
Note that these distributions apply to particle species in equilibrium; hence we must
take into account the interactions that these particles experience. Indeed, this will become
crucial as we discuss the relics of the hot early Universe: the light elements, the microwave
background itself, and the dark matter. If all of the particles in the Universe were in
equilibrium, the abundances of, say, Hydrogen atoms, would be much less than that
12
observed – using m = mp and T = T0 in Eq. 5.14 gives nP ∼ 10−10 cm−3 !
Cosmology 41
5 — Thermodynamics and Particle Physics
If we take the typical case that Γ ∝ T , then, when we also have a ∝ tm (m = 1/2 for
n
(Exercise: Show this.) The average number of interactions experienced by a particle after
time t is therefore less than one if Γ < H, as long as n > 1/m.
By both of these arguments, or by our more detailed discussion of the Boltzmann
equation below, when Γ H, each particle will experience interactions, and can plau-
sibly remain in thermal equilibrium. However, if H Γ, the Universe expands away
more rapidly than the particles can interact, and most particles do not experience the
interaction, and hence fall out of equilibrium. We call this departure from equilibrium
freeze-out, as the comoving abundance of a species that has Γ H is frozen (in the
absence of any further interactions, of course). Note that a species can freeze out either
when it is behaving like radiation (kT mc2 ) or matter (kT mc2 ). In the former
(relativistic freeze-out) we say that the outcome is a hot relic, and in the latter (massive
freeze-out) a cold relic.
42 J. R. Pritchard
5.4.: Relic Abundances
kT (GeV) t (s)
1019 10−43 The Planck Era: Classical GR breaks down
1014–16 10−(35–38) Thermal Equilibrium established; GUT transition? (MX ∼ 1015 GeV)
102 10−11 Electroweak phase transition (MW )
0.1–0.5 10−(5–6) Quark confinement (“chiral symmetry breaking”)
? ? Baryogenesis
10−1 10−4 µ+ µ− annihilation (and freeze-out)
? ? Dark Matter interactions freeze out?
10−3 1 ν decoupling
5 × 10−4 3–4 e+ e− annihilation, leaves mainly γ, (ν ν̄) separately in equilibrium
10−4 180 Nucleosynthesis ⇒ He4 , D, T, Li
10−(8–10) 1010–11 Matter Domination
10−(10–11) 1011–13 H recombination (e + p → H + γ)
Universe becomes neutral and transparent
In Table 5.3.2 we show some of the most important “events” in the history of the
Universe. Some terms, such as “chiral symmetry breaking” and “GUT Transition” come
from particle physics and will not be discussed much further here. Others, such as de-
coupling and freeze-out refer to what happens when interactions fall out of equilibrium
and are crucial to the history of the Universe. I have left a few spaces filled with question
marks, because they depend to some extent on the as-yet unknown physics of the dark
matter and of the creation of the present-day matter/antimatter asymmetry.
Cosmology 43
5 — Thermodynamics and Particle Physics
How did the Universe get to filled with clumpy, baryonic, neutral, matter? To put the
question another way: Why is the Universe interesting?
44 J. R. Pritchard
5.4.: Relic Abundances
1. Baryon number violation. This is the most obvious condition: particle physics must
allow interactions which change the number of baryons. So, for example, we must
allow interactions like
X + Ȳ → B (5.18)
where X is some particle, Ȳ some antiparticle (so the left hand side has net baryon
number B = 0), and B represents some state of particles with net baryon number
B 6= 0.
2. CP violation. If CP is not violated, then in quantum field theory for any interaction,
the related interaction with all particles replaced by antiparticles and vice versa will
have exactly the same rate. That is, the rates of the following two interactions must
be different:
X + Ȳ → B
X̄ + Y → B̄ . (5.19)
Any net change in baryon number (allowed by condition 1) will otherwise be can-
celled by a change in antibaryon number.
Cosmology 45
5 — Thermodynamics and Particle Physics
completely sure where to get the correct baryon number and CP violation, whereas the
physics underlying the preference of, say, nuclei over quarks, and neutral atoms over ions,
is very well understood.
Dfi
= C [{fj }] (5.20)
Dt
where Dfi /Dt is the Liouville operator, just the total time derivative of the distribution
function for particle species i, and the quantity C[{fj }] represents the effect of inter-
actions, which can possibly depend on the distribution function of all other species, j.
In general relativity, the d/dt operation can actually be taken to include the effects of
gravity, more or less automatically by correctly accounting for the coordinates that we
use to describe the manifold. Even with Newtonian gravity, it is still easiest to account
for its effects on the left-hand side, through the Liouville operator. Indeed, we already
know how to do this for the FRW metric.
We will still assume an FRW Universe homogeneous and isotropic and any given time.
Hence, we must take into account the expansion of the Universe. Basically, the equation
must account for the fact that, in the absence of interactions, the comoving number
density of particles is conserved. In fact we already know an equation that describes this:
it is just our non-relativistic matter conservation equation (the fluid equation), Eq. 2.19:
dn ȧ
+3 n=0 (5.21)
dt a
where n gives the number density. That is, in this context, the Liouville operator D/Dt =
d/dt + 3ȧ/a. What about the right-hand side, the collision operator C? Obviously, it
depends on the details of the interactions that we’re considering.
In thermal equilibrium, we know that the number density should obey the equilibrium
distributions discussed in Section 5.2.
Consider a two-particle to two-particle interaction (such as annihilation x + x̄ → y + ȳ
or recombination e + p → H + γ). In equilibrium, the number of forward interactions is
the same as inverse interactions. The forward rate will be σ|v|n2 , where σ gives the cross
section and |v| the relative velocity of the particles. Via detailed balance, the inverse
rate must be such that the collision term is zero when n = neq , the number density
46 J. R. Pritchard
5.4.: Relic Abundances
in equilibrium. Hence the inverse rate must be proportional to σ|v|n2eq . Thus we have
heuristically shown that we can write the whole equation as
dn ȧ
+ 3 n = −hσ|v|i n2 − n2eq
(5.22)
dt a
where hσ|v|i is the so-called thermally-averaged value of the cross section times the ve-
locity, and (so if n = neq it is equivalent to having no interactions). We can rewrite this
as 2
ṅ ȧ n n
+3 = −hσ|v|ineq −1 (5.23)
neq a neq n2eq
or 2
ṅ n ȧ n
= −Γ 2
−1 −3 (5.24)
neq neq a neq
where the interaction rate is Γ ≡ hσ|v|ineq . The time or temperature dependence of Γ
depends on the interaction under consideration, so the solution to this equation usually
needs to be calculated numerically. We show an example in Figure 5.3. In general, higher
Γ (higher hσ|v|i) results in later departure from equilibrium (freeze-out). As we noted
above, a particle that freezes out when kT mc2 is called a hot relic, and one that
freezes out when kT mc2 is a cold relic.
n
1
0.01 Increasing XΣ v \
10-4
10-6
10-8
10-10
m c2
10-12
0.1 1 10 100 kT
Figure 5.3: Freeze-out of particle species as a function of interaction strength, hσvi. The
exponentially-decaying curve is the equilibrium abundance; the lines that “peel off” from
equilibrium correspond to increasing values of hσvi for lower values of the final abundance
and the freeze-out temperature (so later freeze-out time).
As an example, we show the numerical solution to this equation for the case where
hσvi ∝ (kT /mc2 )n for mc2 >
∼ 3kT for various values of the cross-section. We see that
Cosmology 47
5 — Thermodynamics and Particle Physics
higher values of the cross section (higher Γ) correspond to later freeze-out times and lower
abundances. Examining the numerical solution also shows that defining the freeze-out
temperature from Γ(TF ) = H(TF ) and the final abundance from nF = neq (TF ) is a very
good approximation.
Their abundance will then be the value of the equilibrium number density, Eq. 5.10,
evaluated at the freeze-out temperature T = TF , and then converted to a comoving
density, which is then constant from then on. Note that the conversion to comoving
density requires multiplying by a3 /a30 = (1 + zf )−3 = T03 /TF3 , which cancels out the TF3
factor in n;3 the comoving density depends on g but nothing else. If these particles have
sufficient mass that they are non-relativistic today we can simply calculate their energy
density by multiplying the number density by the mass. This gives
m
ν
Ων h2 ' 10−2 . (5.26)
1 eV
Since we know that Ω <
∼ 1 today, this translates into a cosmological
P bound on the neutrino
mass, or more precisely, the sum of all the neutrino masses: mν <
∼ 100 eV.
3
There is an additional factor due to any further relativistic particles that have contributed to the
present-day photon abundance and thus T0 but not to TF , as in the discussion of e+ e− annihilation.
48 J. R. Pritchard
5.4.: Relic Abundances
Although the freeze out temperature of neutrinos turns out not to be critically im-
portant for their present day abundance it is straightforward to estimate. Neutrinos
interact via the weak force, which typically gives an interaction cross section σ ∼ G2F T 2
(GF = αW /MW ' 10−5 GeV−2 is the Fermi energy, αW ' 1/30 is the weak fine struc-
ture constant and MW ' 100 GeV is the weak scale). This gives Γ ' nσv ∼ G2F (kT )5
(since relativistic particles have n ∝ T 3 ), so Γ/H ' G2F (kT )3 m−1 3
Pl ' (T /MeV) (mPl
is the Planck mass). This gives Γ/H = 1 at kTF ∼ MeV — neutrinos freeze out near
kT ∼ 1 MeV.4
Cosmology 49
5 — Thermodynamics and Particle Physics
This has the expected dependence that the energy density in WIMPs decreases for larger
interaction cross-section.
This result is extraordinary, although it might not seem it at first glance. We have
assumed that are WIMPs interact via the weak force and so will have a cross-section
σ ∼ G2F T 2 , evaluating this for TF and assuming that v ∼ c gives hσW vi ∼ 10−26 cm3 s−1 .
Which would predict ΩX h2 ∼ 0.1. This is almost exactly the observed abundance of dark
matter today! Remember that supersymmetry was invented to solve particle physics
problems, not cosmological ones, so this could be just a coincidence. Nonetheless, it is
intriguing that without fine tuning a WIMP could produce the required amount of dark
matter, so much so that this result is often described as the “WIMP miracle”.
We have not gone into too much detail about baryogenesis, neutrinos, and WIMPs, but
in the following lectures, we will discuss two important transitions in the early Universe
using these ideas. First, the creation of neutral hydrogen from the ionized plasma of
electrons and protons at about 400,000 years after the big bang, and then the synthesis
of the light elements from free neutrons and protons at about three minutes.
50 J. R. Pritchard
Chapter 6
MRR 5.2
51
6 — Hydrogen Recombination and the Cosmic Microwave Background
at some time in the past, we must have gone from Γ > H to H > Γ more recently, so
this reaction would have frozen out.
Another consequence of these interactions freezing out is that the photons, once tightly
coupled to the ionized plasma (which is opaque due to Thomson scattering, e− γ → e− γ),
are able to stream freely through the now-neutral, and hence transparent, hydrogen gas.
Thus we see these photons as if freed from a cloud; this is the last scattering surface,
and it forms the cosmic microwave background, which looks to us (using microwave
telescopes) as the surface of a cloud: further away (higher redhsifts) is opaque, nearer is
transparent.
Opaque
Transparent
Figure 6.1: The surface of last scattering. Redshift increases outward. At early times,
the photons are tightly coupled to the charged plasma. At later times they stream freely
through the neutral gas.
52 J. R. Pritchard
6.2.: Equilibrium ionization
and a massless Bose-Einstein distribution for the photons. The total number density of
photons is just the integral of the distribution as in Eq. 5.4:
3
kT
nγ = 16πζ(3) , (6.6)
hc
which also gives the overall baryon density, nB = ηnγ , where η = 2.7 × 10−8 Ωb h2 is the
baryon-to-photon ratio. Statistical chemical equilibrium of Eq. 6.1 requires µp + µe− =
µH + µγ = µH , and the degrees-of-freedom factors are gγ = gp = ge = 2, gH = 4.
Combining the expressions for the equilibrium abundances of the protons, electrons,
and hydrogen atoms gives us the formula
−3/2
gH me mp kT
nH = ne np eB/(kT ) (6.7)
ge gp 2πmH h̄2
If we define the ionization fraction,
np np
Xe = = , (6.8)
nB np + nH
which looks complicated but is really just a quadratic equation for the equilibrium ioniza-
tion fraction as a function of temperature T , or equivalently of redshift by T = (1 + z)T0 .
Exercise: derive Eqs. 6.7-6.8 from the equations given in the previous paragraph. We
show Xe (z) in Figure 6.2.
In the figure, we see that ionization fraction goes from one (fully ionized) to zero
(neutral) around redshift 1300-1400. This corresponds to a temperature kT ∼ 0.3eV
Compare this to the redshift at which kT = 13.6 eV, calculated in one of the problems:
z(13.6eV) ∼ 60, 000. Why the large discrepancy? It is the very small value of η: there are
1
We know from our discussion in Section 5.4.1 that Baryon number is not always conserved, but we
assume that this happens much earlier than this epoch.
Cosmology 53
6 — Hydrogen Recombination and the Cosmic Microwave Background
X
1.0
0.8
Wb h2 0.005
0.6
Wb h2 0.02
0.4
Wb h2 0.1
0.2
z
1200 1400 1600 1800 2000
Figure 6.2: The equilibrium ionization fraction, X as a function of redshift, for different
values of the Baryon density, Ωb h2 .
so many photons for every baryon that even the very small number in the exponentially-
suppressed tail of the BE distribution of photons is enough to keep the hydrogen ionized
until much later. Indeed, we could have derived this just by finding the temperature at
which there is approximately a single ionizing photon with hν > 13.6 eV in the photon’s
Bose-Einstein distribution. Exercise: show this.
6.3 Freeze-Out
In the previous section we calculated the equilibrium ionization fraction. However, be-
cause the interaction rate (i.e., the recombination rate) depends upon the density of free
electrons and protons, as more and more neutral hydrogen is formed, the rate of photon
scattering off of electrons goes down and down, and will eventually freeze out.
Consider Thomson scattering,
e− + γ → e− + γ , (6.10)
54 J. R. Pritchard
6.3.: Freeze-Out
and for lower temperatures the interaction is frozen out. Note that this freeze-out occurs
again at zF ∼ 1100 when the equilibrium ionization fraction is X 1 – despite the very
large number of photons per baryon, it takes a very small number of electrons to make
the mean free path much smaller than the Hubble scale, and hence to make the Universe
opaque. However, once this happens, the recombination interaction ep → Hγ also freezes
out, leaving a final ionization fraction X = Xeq (zF ) ∼ 10−4 , which, although small, is
much larger than the exponentially-reduced equilibrium ionization fraction that would
have otherwise been predicted from Section 6.2.
t
sec
1015
Λ, Wb h2 0.02
1014
Λ, Wb h2 0.1
13
10
H -1 , Wm =0.3
1012
H -1 , Wm =1
11
10
z
1100 1200 1300 1400 1500 1600 1700
Figure 6.3: The expansion timescale (H −1 = a/ȧ) and the mean time between interac-
tions, (λ = 1/Γ) for different values of Ωb h2 and Ωm . For redshifts less than the crossing
point of the curves, the photons and baryons/electrons are no longer interacting.
In fact if we calculate this more carefully using the Boltzmann equation formalism
of the previous lectures, we get the same result (as would be expected): freeze-out of
Thomson scattering at kT ∼ 0.25 eV.
It is important to realize that there are three somewhat distinct events:
• The freeze-in of residual ionization at a higher value than equilibrium, i.e., the
freeze-out of the recombination reaction, Eq. 6.1. (This calculation requires more
of the full Boltzmann equation apparatus to do accurately and we will not do it in
detail here.)
Cosmology 55
6 — Hydrogen Recombination and the Cosmic Microwave Background
56 J. R. Pritchard
Chapter 7
Big-Bang Nucleosynthesis
MRR 5.3
This material is covered in several textbooks, including Rowan-Robinson and Liddle,
but there is an especially classic discussion in S. Weinberg, “The First Three Minutes”.
In this chapter, we will discuss the synthesis of the light nuclei from the initial primor-
dial mixture of protons and neutrons (along with the ubiquitous photons) — Big Bang
Nucleosynthesis (BBN). In broad strokes, this is similar to our discussion of Hydrogen
recombination and the formation of the CMB, but with quite a few extra complications.
At early times, protons, neutrons, and light nuclei such as deuterium, tritium and Helium,
are in equilibrium (specifically called “nuclear statistical equilibrium” in this case, just
as we discussed “ionization equilibrium” in the context of the CMB). As the Universe
expands and cools, different nuclear reactions freeze out, leaving us with relic abundances
of the stable nuclei.
BBN has a few extra complications, however. First, we have to track the abundance
of not just one end product (neutral hydrogen) but of several different nuclei. Second,
one of the starting constituents, neutrons, are unstable when not in a nucleus, with a
half-life of about 11 minutes. Finally, several of the possible end-states (light nuclei) have
binding energies that are small or comparable to kT and so freeze-out can be delayed.
With our calculation of hydrogen recombination, we found in Eq. 6.7 that the equi-
librium Hydrogen density is proportional to exp(B/kT ) where B = 13.6eV is the binding
energy of the hydrogen atom. Similarly, the most strongly-bound light nucleus is 4 He,
with binding energy B4 = 28.5 MeV, so we essentially expect most of the nucleons to end
up in Helium in equilibrium, with other species suppressed exponentially in comparison.
Moreover, the lack of any stable nuclei at all of mass 5 or 8 makes it very difficult to get
to higher-mass nuclei beyond Helium. However, we will see that freeze-out of the inter-
actions will mean that not all the species will remain in equilibrium so a more careful
calculation is required.
Indeed, a full calculation of BBN requires the solution of many coupled differential
equations, but the main results can be calculated from simple principles such as those we
have applied to the CMB.
57
7 — Big-Bang Nucleosynthesis
58 J. R. Pritchard
7.2.: Helium production
Cosmology 59
7 — Big-Bang Nucleosynthesis
where now we have related the helium mass fraction to the neutron-proton ratio calculated
above. At np freeze-out TF ' 1 MeV we had nn /np ' 0.17 ' 1/6. By the time the
deuterium bottleneck is broken, however, it is now tBBN ∼ 3 min, which means that a
non-negligible fraction of the neutrons will have decayed into protons, n → p+ e− ν̄e , with
a half-life of 10.5 minutes. Hence
nn
≈ 0.13 (7.16)
np TBBN
and thus
2 × 0.13
Y ' ' 0.24 (7.17)
1 + 0.13
We can also define the mass fraction of other species Xi = Ai ni /nB which can be found
by considering the freeze-out of the reactions we have considered so far, as well as further
reactions that can build higher-mass nuclei, giving final abundances slightly higher than
the equilibrium abundances. Lithium is produced by the chain
4
He + t → 7 Li + γ 3
He + 4 He → 7 Be + γ 7
Be + e− → 7 Li + νe . (7.18)
60 J. R. Pritchard
7.3.: Observations of primordial abundances
however, lithium is very easy to destroy via collisions with protons, so increasing η actually
means decreasing final lithium abundance. Whereas for higher values of η, there is more
3
He around and we can first produce beryllium via the Helium-3 path,
3
He + 4 He → 7 Be + γ , (7.20)
which then inverse β-decays via electron capture to 7 Li. Because beryllium is actually
more strongly bound than lithium, once this pathway is opened the final lithium abun-
dance increases with η.
Furthermore, as we increase η, more and more d and 3 He is burnt into helium the
reactions Eqs. 7.10-7.11, which freeze out later and later, since the reaction rates are
proportional to η.
Although a full calculation requires the simultaneous solution of coupled differential
equations, it is usually straightforward to perform a perturbation analysis to see the
effect of making a small change in the other cosmological parameters. To give a flavor
of this analysis consider the effect of changing the time (or equivalently, temperature,
TBBN ) that BBN occurs. If the expansion was faster, freeze-out, breaking the deuterium
bottleneck, would have occurred earlier. At this earlier time, neutrons would have had
less time to decay, and the neutron-proton ratio would have been higher, and hence the
Helium mass fraction, Y would have been higher. But why would the expansion rate be
higher? One way is to increase the density of radiation in the universe, since H 2 ∝ ρ, so
any (relativistic) species beyond those we know of in the standard model would increase
the Helium abundance — indeed for quite a while observations of the Helium abundance
(and that of other elements) were the strongest constraint on additional light particles,
in particular neutrinos.
Cosmology 61
7 — Big-Bang Nucleosynthesis
356789%>#9?=;7%Ω@!(
/0//) /0/. /0/( /0/!
/0(+
'"#
/0(*
/0()
D&
A /0('
BBB
"
/0(!
./ −!
A$"% &
123
334
./ − '
!"#$"%
&
./ − )
./ − -
)
+C=$"%
&
(
./ − ./
. ( ! ' ) * + , - ./
356789:;8:&<8;89%65;=8%η × ././
Figure 20.1: The abundances of 4 He, D, 3 He, and 7 Li as predicted by the standard
model
Figure 7.1:of Big-Bang
The mass nucleosynthesis [11] − the
fraction of various bands show
species as a the 95% CLofrange.
function Boxes density.
the baryon
indicate the observed light element abundances (smaller boxes: ±2σ statistical
Bands show the 95% confidence range. Boxes indicate observed light element abundances
errors; larger boxes: ±2σ statistical and systematic errors). The narrow vertical
(smaller indicates±2σ
band boxes: the statistical
CMB measure errors;
of thelarger
cosmic baryon±2σ
boxes: statistical
density, and
while the systematic er-
wider
rors).band
Theindicates
narrowthe
vertical band indicates
BBN concordance rangethe CMB
(both at measure
95% CL). of the cosmic baryon density,
while the wider band indicates the concordance range of direct measurements of the light
element abundances. From Fields and Sarkar, in Amsler et al., PL B667, 1 (2008, 2009)
http://pdg.lbl.gov
62 J. R. Pritchard
Chapter 8
Interlude
63
8 — Interlude
M L2 T −2 = [c]a [h̄]b
= La T −a M b L2b T −b
= M b La+2b T −a−b (8.5)
which is equivalent to
b=1 a + 2b = 2 − 2 = −a − b (8.6)
which are three equations in two unknowns and do not have a solution (a similar argument
holds for lengths or times). So, as long as we keep units on G, we still need to choose
some units in which to report physical quantities, and it is traditional in cosmology and
particle physics to use energy units, in particular electron-volts (or keV, MeV, GeV as
appropriate). Note that any physical quantity can be expressed as an energy to some
power this way. For example, we know that we can relate a length λ to a frequency c/λ
and hence to an energy h̄c/λ — so length can be thought of as having units [energy−1 ].
How do we do the conversion explicitly? We need to find a combination of a length, `
with h̄ and c that gives the appropriate units:
64 J. R. Pritchard
8.1.: Natural Units
Cosmology 65
8 — Interlude
where in the second line we have assumed radiation domination, so ρ = ε ∼ T 4 and used
G−1/2 = MPl . This gives the simple formula
or
[T ]−2 = [Λ] (8.21)
so
[Λ] = T −2 , (8.22)
Which means that in our natural units,
2 3H02
ρΛ c = ΩΛ ρcrit = ΩΛ ×
8πG
∼ 6 × 10−10 J m−3 ∼ 10−47 GeV4 ' (10−12 GeV)4 (8.26)
where in the last we use natural [energy] units. Do we have any fundamental theories
that could predict the value of the cosmological constant? When thought of as an energy
66 J. R. Pritchard
8.1.: Natural Units
density, we have seen that the cosmological constant has the unique property that it
does not change with time, despite the expansion of the Universe — it is the energy of
the vacuum. We also know that our best description of particle physics, quantum field
theory, gives a generic prediction for the value of the vacuum energy density: it should be
ρvac ∼ E 4 (in natural units), where E is some energy scale that describes the high-energy
(“ultraviolet” in particle physics parlance) cutoff of our theory. That is, E is the scale at
which our effective quantum field theory breaks down. Unfortunately, the observed value
of EΛ ∼ 10−12 GeV = 0.001 eV is much lower than any such expected cutoffs — it is well
below the energy scale of everyday physics. Indeed, if Λ is from a true quantum-gravity
theory, we would expect E = EPl , in which case we would have
2
1017
1
ΩΛ ∼ 2
∼ (t2Pl H 2 )−1 ∼ ∼ 10122 , (8.27)
GH 10−44
which is much, much, much greater than the observed value of 0.7. Even if the cosmo-
logical constant were due not to quantum gravity but to supersymmetry, we would still
expect ΩΛ ∼ 1060 , not much better.
There are various solutions, all somewhat unsatisfactory, proposed to solve the cos-
mological constant problem. The first is that the “real” cosmological constant is Λ = 0,
but that there is some other physical mechanism — “dark energy” — that can provide
an energy density with equation-of-state w = p/ρ = −1, not a true vacuum energy. In
fact, a scalar field can provide this. Indeed, we will see that inflation is thought to be
driven by such a scalar field, although the energy scales of inflation and dark energy are
so different that no one has been able to come up with a single mechanism for producing
both epochs of w = −1.
Another possibility is that the properties of the vacuum depend upon the details of
the fundamental theory. In string theory, for example, there are a huge number (possibly
10200 or greater!) ways to compactify down to 3+1 dimensions, and the cosmological
constant could be different in each of them. Then, we may need to use something like
the anthropic argument to find the cosmological constant. The basic idea is due to
Weinberg (“Anthropic Bound on the Cosmological Constant”, Phys. Rev. Lett. 59
(22): 2607–2610. doi:10.1103/PhysRevLett.59.2607): generically, the theory predicts a
distribution of Λ, but most of the distribution is with considerably greater values than
observed. However, with much greater values, the Universe would look very different
than it does today. In particular, it seems that it would be very difficult to form any
structures like galaxies (and likely the stars within them) at all, as the Universe would
start exhibiting accelerated expansion before the structures could form. (We shall see
this in more detail when we discuss large-scale structure). Hence, we wouldn’t be here to
observe the cosmological constant if it were much larger. Thus, the prediction is that we
should observe the largest possible value of Λ consistent with structure formation. This
seems to be, roughly, true. (It should be pointed out that many cosmologists are deeply
troubled by anthropic arguments!)
For the rest of these notes, we will usually use h̄ = c = k = 1 natural units.
Cosmology 67
8 — Interlude
(1 + z)2 1
Ωk (z) = 3
= (8.30)
Ωm (1 + z) Ωm (1 + z)
whereas in the radiation-dominated era
(1 + z)2 1
Ωk (z) = 4
= . (8.31)
Ωr (1 + z) Ωr (1 + z)2
In both cases, |Ωk (z)| is an increasing function of time (i.e., a decreasing function of
redshift). No matter how close the Universe is to flat today, it was even closer in the
past. We know that |Ωk | < ∼ 0.1 today. What does this imply for the curvature at some
early epochs?
−4
• At hydrogen recombination, z ∼ 1000, we have Ωk <
∼ 10 ;
• At matter-radiation equality (z ∼ 104 ), we have Ωk < −5
∼ 10 ;
• at nucleosynthesis, (z ∼ 108 ), we have Ωk < −13
∼ 10 ;
and at earlier and earlier epochs (electroweak symmetry breaking, the Planck epoch), the
requirement gets stronger and stronger. So if the Universe is not flat today, it had to
start out remarkably close to — but not quite — flat. This is not a very generic condition
at all.
68 J. R. Pritchard
8.2.: Open Questions in the Big Bang Model
which is defined by the relation between angular size, θ and physical size L: θ = L/dA .
We expect that the largest distance that physics ought to be able to act is the horizon
size,
cH0−1 ∞ dz 0
Z
dH (z) = . (8.33)
1 + z z E(z 0 )
so we can combine these to find the angular size of the horizon, θH = dH /dA . If we
assume a flat Universe, we have
R∞ 0
dH z
dz /E(z 0 )
θH = = z 0
R . (8.34)
dA 0
dz /E(z 0 )
If we further assume a matter-dominated Universe, the integrals are easy, since E(z) =
1/2
Ωm (1 + z)3/2 :
(1 + z)−1/2
θH = −1/2
∼ (1 + z)−1/2 1 for z 1. (8.35)
1 − (1 + z)
In particular, θH ' 1.7◦ at z ' 1000, and this is about right even if we take the details
of the density of matter, radiation and cosmological constant into account.
This means that any two patches of the CMB sky more than a couple of degrees apart
should not have been in causal contact — so how did they get to be the same temperature
to at least one-tenth of one percent? In fact, we don’t need to use the CMB to see the
problem: even at redshifts of a few, different regions of the sky are in different horizon
volumes, and yet we observe the matter density (as seen in the galaxy density) to be
roughly similar.
Another way to see this problem is to consider the scale of a structure in the Universe,
say, that of a galaxy. The physical scale of a galaxy, λgal , grows along with the scale factor,
λgal ∝ a ∝ t2/3 (MD) or λgal ∝ t1/2 (RD). But (in a Universe dominated by matter,
radiation, or curvature) the horizon scale grows as dH ∼ t ∼ H −1 . Today, λgal < dH and
we say that the scale is “inside the horizon” but because they grow at different rates,
at some point in the past we must have had λgal > dH , “outside the horizon”; the time
at which the dH = λgal is called horizon-crossing for that scale (see Figure 8.1). We
expect that structures can only grow when they are inside the horizon (due to causality)
but we must further be able to set up initial conditions so that this can happen. And
Cosmology 69
8 — Interlude
for astrophysically-interesting scales, this happens at a time for which we thing we have
a good understanding of the physics (e.g., the galaxy-size scale of λ ∼ 10 Mpc enters at
approximately the epoch of matter-radiation equality).
dH∝t
log length
λgal
tHor(λ)
log time
Figure 8.1: Physical scales inside and outside the horizon. Prior to (i.e., to the left of)
horizon-crossing at tHor (λ), the scale is outside the horizon; afterwards it is inside the
horizon and causal physics can act.
Note that we could have equally made this argument in terms of comoving scales,
lgal = λgal /a = const, which now must be contrasted with the comoving horizon distance
χH = dH /a. In a universe with matter, radiation and curvature, χH ∼ (aH)−1 and the
comoving Hubble length always grows with time: the horizon grows to encompass larger
and larger scales, so any scale now inside the horizon was once outside.
70 J. R. Pritchard
8.2.: Open Questions in the Big Bang Model
We can work out Ωmon by dividing by ρcrit = 3H02 /(8πG) ∼ H02 m2Pl :
3
T 3 T 3 mmon
mmon TGUT
Ωmon ∼ GUT 5 0 2 ∼ 1011 16 (8.37)
mPl H0 10 GeV 1014 GeV
where we have used H0 ∼ 10−42 GeV, T0 ' 2 × 10−13 GeV, mPl ' 1.2 × 1019 GeV and
put in typical values for the GUT scale and the monopole mass.
Since we know that Ω ∼ 1 we know that this is not possible. (It is often said that these
relics would “overclose” the Universe, but a better interpretation would be to say that
such a density at this early time would have caused the Universe to be closed and collapse
again on a very short timescales, rather than continue expanding for the 14 billion years
— and counting! — that we observe.) Moreover, our observations of, say, the light-
element abundances also imply that the Universe must have been radiation-dominated
at least prior to about three minutes after the big bang.
This problem was thought to be such an important issue that it was the primary
motivation for Guth’s original model of inflation — it is often referred to more specifically
as the monopole problem. In fact, we do not know if a GUT transition actually
happened in the early Universe, but there are various other transitions that may have
happened, resulting in a high density of very massive particles.
2
Not to be confused with the superstrings of string theory!
Cosmology 71
8 — Interlude
72 J. R. Pritchard
Chapter 9
Inflation
Liddle Ch. 13
In the last chapter, we discussed the way in which the Universe in which we live
started out in a very special state: nearly flat, nearly homogeneous, and dominated by
radiation. It is of course possible that these initial conditions are just a raw fact that we
have to learn to deal with, but we would prefer to find a causal mechanism to enforce these
conditions. In about 1980, Alan Guth in the USA (and independently Alexei Starobinsky
and Andrei Linde in the former USSR) came up with a mechanism that takes a broad
range of initial conditions and makes them all look like a flat, homogeneous, radiation-
dominated Universe — inflation. Furthermore, it was quickly realized that inflation also
provided a mechanism to generate density fluctuations of just the right character to grow
into the large-scale structure we observe in today’s distribution of galaxies, as well as in
the fluctuations in the CMB which were first observed in the early 1990s.
73
9 — Inflation
As we have seen, usually this quantity behaves like (“is approximately proportional to”)
the comoving horizon distance 1/(aH), and both of these are increasing functions of time.
But for a fixed present day value 1/(a0 H0 ) we can make χH as large as we want if we can
make 1/(aH) decrease with time — because that makes it increase back towards a = 0
and we can therefore increase the value of the integral. If we can do this then the Hubble
scale grows more slowly than any fixed comoving distance, as we show in a cartoon in
Figure 9.1. After an accelerating expansion, the Hubble length has shrunk with respect
to the comoving coordinates.
Hubble Volume
Inflation
Comoving "grid"
Figure 9.1: The Universe before and after inflation. The gridlines represents the comoving
coordinate system, and the circle the Hubble volume 1/H, which has shrunk with respect
to the comoving coordinates after inflation.
We see how this solves the Horizon problem in Figure 9.2. The accelerating expansion
means that the Hubble scale remains constant, but comoving scales increase much more
rapidly. Note that in this case the true horizon scale is now much larger than the Hubble
length.
74 J. R. Pritchard
9.1.: Accelerated Expansion
dH~1/H∝t
duration of inflation
λgal∝a
1/H∝const
log length
λgal∝eHt
log time
Figure 9.2: Scales entering and leaving the Hubble scale, which is the apparent (but not
actual) horizon in a Universe with a period of accelerating expansion.
Thus, accelerating expansion should be able to solve the horizon problem. It is clear
that it can also solve the flatness problem as it means that the value of |Ωk | = |k|/(aH)
gets driven closer and closer to zero while acceleration is occuring. Heuristically, this
makes sense: if we rapidly expand a curved surface (relative to our coordinate system) it
looks more and more flat in those coordinates.
Finally, acceleration solves the relic (monopole) problem in much the same way: it
dilutes the number of massive relic particles in a given (physical) volume much faster
than “ordinary” decelerating expansion. In fact, it is a little more complicated than
this, because we still have to find a way to stop the accelerated expansion and make the
Universe radiation-dominated after the period of accelerated expansion — this is called
reheating.
After we discuss what kind of matter is necessary in order to make the Universe
accelerate, we will return to these issues and calculate just how long inflation needs
to last. But the idea of inflation is very simple: starting from a fairly generic state,
the Universe undergoes accelerated expansion, which increases the scale factor by many
orders of magnitude. However, this cools the Universe down proportional to the scale
factor, so we must find a mechanism for reheating: stopping the accelerated expansion
and converting the energy density of the Universe into radiation, at which point the
Universe looks like a hot, radiation-dominated, flat big-bang model. We show these steps
in Figure 9.3.
Cosmology 75
9 — Inflation
Inflation
a(t)
Reheating
T(t)
time
Figure 9.3: The scale factor and temperature in inflation and reheating, leading to a
Universe that looks like a standard radiation-dominated hot big bang.
76 J. R. Pritchard
9.1.: Accelerated Expansion
For this w = −1 case, we have already seen how it affects the expansion, back in
Chapter 3. It gives a de Sitter Universe with exponential expansion:
hp i hp i
a ∝ exp Λeff /3 t ∝ exp 8πGρinf /3 t ∝ eHt , (9.5)
Cosmology 77
9 — Inflation
where we have set N = Hδt as the “number of e-folds” that inflation lasts. We have
plentiful evidence that the universe is isotropic at or near the current Hubble length, so
we need
d > H0−1
2
3 K TRH
eN > H0−1
TRH mPl
3 K TRH
> (9.8)
H0 mPl
This is what is needed to solve the horizon problem. In order to solve the flatness
problem we can make a similar analysis. Before inflation, we start with some initial
value of the curvature parameter, Ωk (ti ) = −k/(ai Hi )2 . Today, the scale factor is a0 =
ai eN TRH /T0 , accounting for inflationary and ordinary expansion. So the current curvature
2
−k −k −k
−N T0 Hi
Ωk (t0 ) = = = e
(a0 H0 )2 (H0 ai eN TRH /T0 )2 (ai Hi )2 TRH H0
2
T0 TRH
= Ωk (ti ) e−N = Ωk (ti )e2Nmin −2N (9.10)
H0 mPl
where Nmin is the minimum number of e-folds required to solve the horizon problem,
Eq. 9.9. Thus, if Ωk (ti ) is of order one, the same amount of inflation that solves the
horizon problem will solve the flatness problem; increasing the initial curvature parameter
by orders of magnitude requires only increasing the number of e-folds logarithmically. In
fact, realistic inflation models tend to last much longer than Nmin , so this is easily satisfied.
Finally, inflation decreases the relic abundance down to acceptable levels from Ωmon (t0 ) ∼
1011 (Eq. 8.37). The monopoles are produced at TGUT with number density n(tGUT ) ∼
3
HGUT which is diluted by a factor (Ti /TGUT )3 by the beginning of inflation, a factor of
exp(−3N ) during inflation, and subsequently by another factor of (T0 /TRH )3 . All to-
gether, this is decreased by a factor of exp(−3N )Ti3 /TRH 3
from the no-inflation prediction
of Eq. 8.37. We must have Ti > TRH (or otherwise you would inflate again after reheat-
ing), but typically the two are comparable, so the main effect is from the exponential.
We need 3N > ∼ 11 ln 10 ∼ 25, which is easily satisfied if we already solve the horizon and
flatness problems.
Thus, we see that generically we need something like 60 e-folds to solve the various
problems. To put this another way, with at least that much expansion, very generic initial
conditions (inhomogeneous, curved, lots of heavy particles) are funnelled into what seems
to be a very special state: smooth, flat, and radiation-dominated.
78 J. R. Pritchard
9.2.: Inflation via a scalar field
1
L(ϕ) = ∂ µ ϕ∂µ ϕ − V (ϕ) (9.11)
2
and the stress-energy tensor is then given by
T µν = ∂ µ ϕ∂ ν ϕ − Lg µν (9.12)
where g µν is the metric. For comparison, recall that we wrote down the stress-energy
tensor of a perfect fluid with pressure p and density ρ:
If we assume that the field is spatially homogeneous, in fact the stress-energy takes on
the perfect fluid form, with
1
ρϕ = ϕ̇2 + V (ϕ)
2
1
pϕ = ϕ̇2 − V (ϕ) ; (9.14)
2
the correction terms for spatial inhomogeneities are of order ∇ϕ2 /a2 where ∇ is the
comoving-coordinate gradient. Because of the 1/a2 factor, the effect of any spatial gra-
dient is quickly rendered irrelevant by exponential accelerated expansion.
We can derive the equation
R 4 of√motion in several ways. For a general scalar field, we
could vary the action, S = d x −gL, where g is the determinant of the metric tensor
or we could use the covariant conservation of the stress-energy tensor ∇ν T µν = 0. But
both of these require a bit more relativity than we’ve developed here. Instead, we can
just plug in these expressions for the pressure and density into the fluid equation
ρ̇ + 3H(ρ + p) = 0 (9.15)
which gives
ϕ̈ + 3H ϕ̇ + V 0 (ϕ) = 0 . (9.16)
If it were not for the friction term proportional to H (“Hubble drag”) note that this
is exactly the same as the equation of motion for the position of a particle in a one-
dimensional potential V , so we can use our intuition from that situation.
Cosmology 79
9 — Inflation
During inflation, we are looking for solutions that look like p ≈ −ρ, which means
that the time-derivative terms must be negligible, so ρ ≈ +V , p ≈ −V . In order to do
this, the velocity ϕ̇ must be negligible — the potential must be very flat. However, for
inflation to end, the potential cannot be completely flat — it must eventually fall into a
potential well where it can oscillate, which corresponds to a mass in quantum mechanics
— a term in the potential like
1
V (ϕ) = m2ϕ ϕ2 + · · · (9.17)
2
gives a mass, so nearly any concave potential gives mass to the field. Hence we expect
that a potential of the form given in Figure 9.4 will have the right properties.2 In fact
we further expect that the ϕ field should couple to other particles, which would induce
another “friction” term Γϕ on the left-hand side of the equations of motion, where Γ
represents the decay rate of ϕ into other particles. If these particles are light, we end up
with a radiation-dominated Universe.
When the field is sitting on the approximately flat part of the potential, we say that
it is in the “slow-roll” regime, with V 0 (ϕ) very small.
Slow-roll
(acceleration)
V(!)
during inflation:
-pϕ=ρϕ≈V
Oscillations
! (reheating)
Figure 9.4: The inflationary potential. The slow-roll regime is when the potential is
relatively flat, and the reheating regime is when it strongly curved.
80 J. R. Pritchard
9.2.: Inflation via a scalar field
were the initial fluctuations which have subsequently grown into the large-scale structure
we observe in the Universe today generated?
The basic mechanism is very simple, and is a relatively generic consequence of the com-
bination of inflation’s accelerated expansion with the randomness of quantum mechanics.
Any system will exhibit quantum fluctuations on very small scales. Inflation takes these
fluctuations — initially on very small scales — and blows them up very rapidly. Once
they are larger than the apparent horizon (the Hubble length, H −1 ∼ ct), they are frozen
in and behave as completely classical fluctuations. (Even though the actual horizon is
much larger, the behavior of fluctuations at any given time is still controlled by the speed
of light.) Outside the horizon, fluctuations can only evolve in a very simple way, due
to the finite speed of light, so only when a scale re-enters the Horizon do overdensities
(lumps) begin to collapse.
Without quantum field theory it is difficult to do the calculation precisely, but we can
at least make a plausible argument using dimensional analysis. We start with our scalar
field, which we will split into
ϕ = ϕcl + δϕQM (9.18)
where “cl” labels the classical evolution, and “QM” are the quantum fluctuations. We
wrote down the Lagarangian density for our scalar field, L = ∂ µ ϕ∂µ ϕ/2−V (ϕ) (Eq. 9.11),
so the units on the scalar field are
If we are really in the slow-roll regime, then there is only one quantity with these units
in our problem, and that is the (exponential) expansion rate, H = const. Since we know
that the average of our quantum fluctuations should be zero, this implies that we can use
this to fix the variance of our fluctuations, hδϕ2 i ∼ H 2 . In fact, a more careful analysis
gives 2
3 2 2 H
k Pϕ (k)/(2π ) = hδϕ ik ' . (9.20)
2π
In this equation, Pϕ (k) is the power spectrum of the ϕ field at spatial frequency k,
using Fourier-transform conventions to be defined in the next chapter. The factor of
2π on the right-hand side arises because this is actually the so-called Gibbons-Hawking
temperature associated with the horizon in a de Sitter Universe (the equivalent of the
Hawking temperature associated with the horizon of a black hole), TGH = H/(2π).
Because H is approximately constant during inflation, we can integrate this expression
to get the total mean-square fluctuation in ϕ, integrated over all frequencies (which are
sometimes called “modes”), which ends up giving
2
2 H
hδϕ ik ' N × . (9.21)
2π
where N is the number of e-folds of inflation from before. In the limit of exactly expo-
nential (de Sitter) expansion, this would result in an initial power spectrum of density
fluctuations (which we will define more precisely in the next chapter) with the so-called
Cosmology 81
9 — Inflation
P (k) ∝ k ns , (9.22)
where ns is the scalar spectral index (another name for density perturbations are scalar
perturbations, corresponding to actual fluctuations in the curvature of the spacetime
manifold, described by a scalar number), and we usually have ns just below one, with
details depending on V (ϕ). The amplitude of the spectrum depends on the coupling
constants present in V (ϕ), and are constrained to be quite small in most models by the
relatively small amplitude of temperature and density fluctuations observed in the CMB
and on large scales today.
We will see in the next chapter that it is straightforward to understand the evolution
of the power spectrum of such fluctuations once they exist.
Moreover, for a weakly-coupled scalar field (and we have just noted that observations
seem to require weak coupling), the distribution of these fluctuations will be very close to
Gaussian (in field theory, a free field is exactly Gaussian). Hence, once we have described
these second moments, there is no more information available to us.
In a similar manner, inflation also creates a background of gravitational radiation
(gravitons, or “tensors”). Gravitational radiation does not directly create lumps and
voids (it does not couple directly to the density of matter) but the movements it induces
are indeed visible, although as yet undetected, as patterns in the polarization of the CMB.
The gravitational radiation is described by a separate power spectrum,
PT (k) ∝ k nt (9.23)
where now nt is usually just below zero, and the amplitude is governed by the value of
V (ϕ) during inflation, i.e., by the energy scale of inflation.
The observation of these tensor modes via CMB polarization is one of the main goals
of the next generation of CMB experiments (beyond the Planck Satellite).
82 J. R. Pritchard
Chapter 10
Structure Formation
Schneider Ch. 7
83
10 — Structure Formation
chronous gauge in which all locations use the proper time of an observer located their as
the time coordinate. Another is the Newtonian gauge in which the equation of motion
for the perturbations looks most like the Poisson equation of nonrelativistic gravity.
It turns out that well inside the apparent horizon, well away from regions of very strong
gravity such as black holes, and at speeds considerably lower than c, most reasonable
choices of the gauge agree. It is mostly outside the horizon that the choice matters.
However, we shall be careful to use only physical “gauge-invariant” descriptions on such
large scales and not use equations that refer to specific coordinate systems.
It is clear that our understanding of any physical theory will only be such that it
will predict the statistics of the density perturbation, rather than the actual values as a
function of our time and space coordinates. Hence, we need to consider suitable averages
of δ and functions of δ.
Given our definition of δ(x, t) it is obvious that the spatial average at a given time
requires
hδ(x, t)i = 0 . (10.2)
(Actually, there is a subtlety, as there are three different averages that we could con-
template here. The first is a over all space at a given time; the second is an average
with respect to the probability distribution of δ. (The equivalence of these two is math-
ematically similar to ergodic theory in thermodynamics.) Of course, neither of these two
average can be performed in the real Universe, in which case averages are over observ-
able parts of the Universe. We shall assume no distinction between these possibilities in
practice.)
The next moment, the variance, will be nonzero:
We have defined ξ, the density correlation function, which is also the second moment of
the density distribution. It is only a function of the distance between the points x and
y, which is a statement about the statistical isotropy of the Universe. Our underlying
description of the Universe should be invariant if we shift the origin and orientation.
(This is the spatial equivalent to a stationary random process in time.)
The correlation structure is much simpler if we consider the Fourier Transform of δ
(as are the equations of motion, as we will see soon):
Z
δ̃(k, t) = d3 xeik·x δ(x, t) (10.4)
With these definitions, we can calculate the correlation function of our Fourier-Transformed
84 J. R. Pritchard
10.1.: Notation and Preliminaries
Z Z
0 0
= d x d3 x0 ei(k·x+k ·x ) ξ (|x − x0 |)
3
Z Z
3 0 0 0
= dy d3 x0 ei(k·y+k·x +k ·x ) ξ(y)
Z Z
3 0 i(k+k0 )·x0
= dx e d3 y eik·y ξ(y)
We can also calculate the statistics of some more physically relevant quantities. Con-
sider the density fluctuation measured not at a point, but in spheres of some radius R.
At any point, we can measure the mass in that sphere
Z Z
δMR 1
(x) = 3
d y δ(x + y) = d3 u WR (u − x)δ(u)
3
(10.10)
MR 4πR /3 y<R
Cosmology 85
10 — Structure Formation
where in the second equality we define the window function, WR (x), which is 1/V within
the volume, and 0 outside. We can now calculate the second moments of the MR distri-
bution from the second moments of δ. First note that Eq. 10.10 is a convolution, so in
Fourier space the transform of δMR /MR is just the product of the transforms of δ and
the window function. Hence
d3 k ik·x
Z
δMR
(x) = e δ̃(k)W̃R (k) (10.11)
MR (2π)3
with
Z
1
W̃R (k) = d3 x eik·x WR (x)
4πR3 /3
Z
1
= 3
dx x2 dφ d cos θ eikx cos θ
4πR /3 x<R
3 sin kR
= − cos kR
(kR)2 kR
≡ W̃ (kR) (10.12)
(in particular, W̃ (0) = 1, so the zero-radius sphere is equivalent to just using δ itself), so
* 2 +
d3 k d3 k 0
Z
δM 2
0 0
= δ̃(k )W̃R (k )
M (2π)3 (2π)3
R
d3 k 2 k3
Z 2
Z
dk
= W̃ (kR) P (k) = W̃ (kR) P (k)
(2π)3 k 2π 2
Z 2
dk
= W̃ (kR) ∆2 (k) (10.13)
k
where we define ∆2 (k) = k 3 P (k)/(2π 2 ), which is the contribution per logarithmic integral
to the mean-square fluctuation, and which has no units. (Recall that we encountered the
similar combination k 3 Pϕ (k)/(2π 2 ) related to the inflaton field in Eq. 9.20). Roughly
speaking, ∆2 (k) gives the density fluctuation on a length scale L ∼ 2π/k.
86 J. R. Pritchard
10.2.: Spherical Collapse
a higher Ω. But our mean density is Ω = 1, and hence our overdense sphere will behave
like an Ωm > 1 Universe — it will expand and recollapse.
Outside of our carved-out sphere, r > r2 we just have the usual equations (with an
overbar to represent the mean):
2
ā˙ 8πG
= ρ̄ (10.14)
ā 3
ρmax 9π 2 9π 2
= ' ' 5.55 . (10.19)
ρ̄(tmax ) 16 16
This is equivalent to δmax = 4.55. If instead we concentrate not on the full collapse,
but just on the very earliest times after ti , we find that δ ∝ a(t) ∝ t2/3 . These
small overdensities are called the “linear regime” for reasons that will be more obvi-
ous in the next section. We can usually calculate things much more readily in the lin-
ear regime, and so it is common to compare this nonlinear (but idealized) case with
what would happen if linear evolution continued. Between ti and tmax , we would have
δLin /δi = āmax /āi = (tmax /ti )2/3 = (3/5) (3π/4)2/3 (amax /ai ) ' 1.06/δi . So, irrespective of
the starting conditions, maximum density (a.k.a. “turnaround”) occurs when the linear
density contrast would have been δ ' 1.06
In a perfectly uniform universe, this overdensity would just collapse down to a point.
But in a more realistic scenario, it will virialize and convert its gravitational energy into
random kinetic energy (by a process known as “violent relaxation”!). At the maximum,
Cosmology 87
10 — Structure Formation
with KE = 0, the overdensity has total energy P E = Egrav ' −3GM/(10rmax ). After
virialization to form a sphere of radius rvir , we have KE = −P E/2 and therefore a
total energy Egrav /2 ' −3GM/(20rvir ). For the energy to remain constant, we need
rvir ' rmax /2 or ρvir ' ρmax /8. Symmetry of the solutions for a closed universe mean
that collapse occurs at tvir = 2tmax , so that the background density at this time is just
ρ̄(tvir ) = ρ̄(tmax )/4. Combining these results gives
ρ(tvir ) 8ρ(tmax ) 9π 2
1 + δvir = = = 16 × = 9π 2 ≈ 178. (10.20)
ρ̄(tvir ) ρ̄(tmax )/4 16
A similar calculation to before gives a linear density contrast δL (tvir ) = 1.68. These values
provide key insights into structure formation. We would expect a collapsed structure to
have a density ∼ 178 the density at the time of collapse and indeed this appears to hold
true for structures like galaxies and clusters of galaxies we observe.
∂v ∇P
+ (v · ∇)v = − − ∇Φ Euler . (10.23)
∂t ρ
88 J. R. Pritchard
10.3.: Linear Perturbations
We also need the Poisson equation relating gravitational potential and matter density,
and an equation of state linking the pressure and density, usually given in the form of an
expression for the [adiabatic] speed of sound,
2 ∂p
cs = equation of state . (10.25)
∂ρ adi
These equations are too complicated to solve in complete generality. Instead, we will
use perturbation theory. It is easy to see that a zeroth-order solution is given by v = 0,
ρ = ρ0 = const, p = p0 = const (assuming no spatial variation in the equation of state)
and ∇Φ0 = 01 We will the write all of our variables as
ρ = ρ0 + ρ1 p = p0 + p1 v = v1 Φ = Φ 0 + Φ1 (10.26)
where 0 refers to the unperturbed solution, and 1 to our small perturbations. To linear
order (i.e., any products of first-order quantities, such as ρ1 × v1 , are neglected), and
subtracting off the zeroth-order solution, our equations become
∂ρ1
+ ρ0 ∇ · v1 = 0 (10.27a)
∂t
∂v1 1
+ c2s ∇ρ1 + ∇Φ1 = 0 (10.27b)
∂t ρ0
2
∇ Φ1 = 4πGρ1 (10.27c)
If we take the divergence of (b) and substitute in (c) and the time derivative of (a), we
get a single, second-order differential equation:
∂ 2ρ
− c2s ∇2 ρ1 = 4πGρ0 ρ1 . (10.28)
∂t2
This equation is straightforward to solve by considering plane-wave solutions (or, equiv-
alently, writing these equations in terms of the spatial Fourier transform of ρ). In terms
of our density perturbation δ = ρ1 /ρ0 , we will try
−ω 2 ρ1 + k 2 c2s ρ1 = 4πGρ0 ρ1
−ω 2 + k 2 c2s = 4πGρ0 (10.30)
1
Actually, this isn’t true, as it contradicts the Poisson equation! This is sometimes called the Jeans
swindle, and it is nonetheless a good approximation for what actually goes on. Moreover, the same issues
do not arise in the expanding case we will discuss next.
Cosmology 89
10 — Structure Formation
If k > kJ , ω 2 > 0 and the solution is oscillatory; if k < kJ , ω 2 < 0 and the solution is
exponentially growing or decaying (there is usually one of each as this is a second-order
equation). In the limit k kJ , ρ ∝ e±t/τ with timescale τ ' (4πGρ)−1/2 .
Basically, small-scale perturbations oscillate as sound waves, whereas large-scale per-
turbations can grow. Another way to see how the behavior changes is to compare the
gravitational timescale τg ∼ (Gρ)−1/2 with the pressure timescale (τp ∼ λ/cs ' (kcs )−1 ).
When τg < τp , collapse has time to occur before pressure can act to respond. We will see
very similar behavior in the case of the expanding universe.
Note that we have found another reason why our Fourier analysis of Section 10.1 is
valuable: individual Fourier modes evolve independently.
(I am ignoring the fact that we have not actually solved for the velocity field, and
in fact there is some subtlety: from Eq. 10.27a, the density only allows us to solve for
∇·v, which leaves ∇×v undetermined — this is just conservation of angular momentum,
which, to linear order, doesn’t couple to the density field.)
• The continuity, Euler, and Poisson equations must be updated to take account of
the possible presence of relativistic matter (e.g., radiation); and
In fact, we will simplify a bit from the most complex possible situation of multiple com-
ponents (e.g., matter plus radiation), each possibly with its own fluctuations. Instead,
we will largely concern ourselves with the fluctuations in the (NR) matter component,
even when the Universe is radiation dominated.
We have already seen how to modify our equations to account for relativistic matter.
In the continuity equation, we replace ρ with ρ + p = ρ(1 + w) as in the homogeneous
fluid equation, Eq. 2.42:
ρ̇ + (1 + w)∇ · (ρv) = 0 . (10.33)
90 J. R. Pritchard
10.3.: Linear Perturbations
In this case, the density and pressure refer to the total from all components. Each
component will also satisfy its own separate conservation equation (e.g., the number of
dark matter particles is conserved as long as we are long after freeze-out for them).
Similarly, we must replace ρ with ρ + 3p in the Poisson equation, as in the second-
order Friedmann equation Eq. 2.41 (which is, after all, an equation for the acceleration,
so it should not be surprising that is this combination that is the relativistic source for
gravitational acceleration):
∇2 Φ = 4πGρ(1 + 3w) . (10.34)
In the expanding universe, we will write our perturbations as ρ = ρ̄(1 + δ), p = p0 + δp =
ρ̄(w + c2s δ), v = v0 + u = Hx + u and Φ = Φ0 + φ.
We also need to change from fixed coordinates x to comoving (also called Lagarangian)
coordinates q. These are related by x = aq, with the unperturbed velocity v0 = Hx.
We must calculate the time derivative of some function f (x, t),
∂f ∂f (aq, t)
=
∂t q ∂t q
∂f X ∂f d(aqi )
= +
∂t x i
∂xi dt
∂f
= + (v0 · ∇x ) f (10.35)
∂t x
and the spatial derivatives are related by a simple change of variables,
1
∇x f = ∇q f . (10.36)
a
We also need the specific cases ∇x · x = 3, ∇x · v0 = 3H and (u · ∇x )v0 = Hu.
With all of these developments, we can derive the equations for the NR matter per-
turbation in comoving coordinates, again to linear order,
1
δ̇ + ∇ · u = 0 (10.37a)
a
ȧ c2 1
u̇ + u + s ∇δ + ∇φ = 0 (10.37b)
a a a
2 2
∇ φ = 4πGa ρ̄m δ , (10.37c)
which can be compared to the non-expanding case, Eq. 10.27. Note that in these equations
δ = δρm /ρm where ρm is the average NR matter density (not the total density). We allow
for the possibility of a non-zero sound speed for the matter (e.g., for baryons).
We can again combine these to form a single second-order differential equation
ȧ c2
δ̈ + 2 δ̇ − s2 ∇2 δ − 4πGρ̄m δ = 0 . (10.38)
a a
Compared to the non-expanding, matter-only case, Eq. 10.28, this equation has a few
new features. First, note the explicit presence of the factors of w from the continuity and
Cosmology 91
10 — Structure Formation
Poisson equations. Second, there is a factor of 1/a2 in the spatial gradient term. Finally,
there is a new term, 2Hδ, occasionally referred to as “Hubble Drag”, which will serve to
slow the growth of perturbations compared to the exponential form in the non-expanding
case.
Nonetheless, we can analyze this equation in much the same manner as before. By
substituting in a plane-wave solution (or, equivalently, by Fourier transforming δ), we
transform ∇2 → −k 2 and get
2
ȧ 2 k 4πGρ̄
δ̈ + 2 δ̇ + cs 2 − δ=0. (10.39)
a a c2s
where δ0 gives the spatial part of the solution at some particular time (i.e., the initial
conditions), and D(t) is known as the growth factor. In the matter-dominated epoch, the
equation for D is just
2
ȧ 3 ȧ
D̈ + 2 Ḋ = 4πGρ̄D = D, (10.42)
a 2 a
where we have used the first-order Friedmann equation to eliminate ρ̄. This is a second-
order equation, hence with two solutions. Often, one of these will grow in time, and the
other will decay, and these are known, respectively, as the growing mode and decaying
92 J. R. Pritchard
10.3.: Linear Perturbations
mode. In both matter- and radiation-dominated Universes, this has power law solutions.
First, matter-domination, with w = 0, a ∝ t2/3 , ȧ/a = 2/(3t), has
D = C1 t2/3 + C2 t−1 . (10.43)
2/3 −1
There is a growing mode D+ ∝ t ∝ a and a decaying mode D− ∝ t ∝ H. Since the
Jeans length is infinite, this growth applies on all scales when the pressure is negligible.
The more general equation, valid during radiation domination and on superhorizon
scales (which can nonetheless be derived from the above Newtonian equations with the
appropriate special-relativistic generalizations to include the effects of pressure), is
2
ȧ 3 ȧ
D̈ + 2 Ḋ = (1 + w)(1 + 3w)D , (10.44)
a 2 a
In a radiation-dominated universe, now with w = 1/3, a ∝ t1/2 , ȧ/a = 1/(2t), so the
solution to this equation is
D = C1 t + C2 t−1 , (10.45)
with growing mode D+ ∝ t ∝ a2 and a decaying mode D− ∝ t−1 ∝ H. We saw that the
Jeans length in an RD universe is comparable to the Hubble scale, so only perturbations
outside the (apparent) horizon grow in this way.
where the subscript labels the species and the sum is over all species, and δj = δρj /ρj
where ρj is the average of particle species j. (As written, this equation ignores the
factors of w since we will use it below to calculate the evolution of the w = c2s = 0 matter
perturbations.)
In particular, we can use this for the common case of matter perturbations in a
radiation-dominated universe. In this P case, we care about i = m = matter, so cm = 0
and we can ignore the 4πGρ term since j ρj δj = ρm δm +ρr δr ; the first term is suppressed
since ρm ρr and the second since δ ' 0 since (as we saw above) there are only sound
waves, not growing perturbations, to the radiation itself. Hence, the differential equation
becomes
ȧ
0 = δ̈m + 2 δ̇m
a
1
= δ̈ + δ̇ . (10.47)
t
Cosmology 93
10 — Structure Formation
since there are no spatial derivatives, we can consider this as an equation for the growth
factor, with solution
Dm = C1 + C2 ln t RD . (10.48)
The growing mode is no longer a power law, but logarithmic. This is sufficiently slowly
that, essentially, matter perturbations do not grow during radiation-dominated periods.
We can similarly calculate the behavior of matter perturbations in an open curvature-
dominated universe. We now go back to our original single-component Eq. 10.42, with
w = c2s = 0, but use the appropriate a ∝ t. Now, the term 4πGρ̄ = (3/2)Ωtot (t)H 2 , so
2 3 2
0 = δ̈ + δ̇ + 2 Ω(t) ' δ̈ + δ̇ (10.49)
t 2t t
where we ignore the last term since not only do we eventually have Ω(t) 1, but
moreover it is falling with time so it must eventually become negligible. The solution is
Dm = C1 + C2 /t CD . (10.50)
Again, there is no growing mode.
Finally, we consider the behavior of matter perturbations in a universe dominated by
a cosmological constant: w = −1, a = exp(Ht), ȧ/a = H = const:
0 = δ̈ + 2H δ̇ − 4πGρ̄δ ' δ̈ + 2H δ̇ , (10.51)
where again, we find that we can ignore the final term, this time since ρ̄ is quickly diluted
by the exponential expansion, and the solution is
Dm = C1 + C2 e−2Ht ΛD . (10.52)
Yet again, we find that perturbations do not grow.
94 J. R. Pritchard
10.4.: The processed power spectrum of density perturbations
2 D2 (t)
δH (t) = ∆2 (k = aH, t) = ∆i (k = aH) = const . (10.53)
D2 (ti )
Here, k = aH is the comoving wavenumber corresponding to the comoving length scale
∼ 1/(aH) at time t, D(t) is the growth factor calculated in Section 10.3.2, and ti is some
suitably early time (just after inflation, say). Because this combination doesn’t change
with time, the behavior is said to be scale-free (which is a slightly different and more
restrictive use than in other fields of physics).
Using the results of the previous section, we can show that these two definitions are
equivalent. In a radiation-dominated Universe, for scales outside the horizon, we have
D ∝ a2 , a ∝ t1/2 , and aH = ȧ ∝ t−1/2 ∝ a−1 so
2
δH ∝ a4 (aH)3 (aH)ns ∝ a4 × a−3−ns
∝ a1−ns (10.54)
Cosmology 95
10 — Structure Formation
where ti is some early initial time (i.e., when P (k) = Ak ns is a good description over all
scales). Ti2 (k) is the transfer function which, for linear evolution in Fourier space, acts
separately on each scale.
We need to combine information for scales inside and outside the horizon, before and
after matter-radiation equality. We summarize those results in Table 10.4.2.
Inside Outside
RD - a2
MD a a
Table 10.1: The growth of perturbations, inside and outside the apparent horizon during
matter- and radiation-domination. The entries correspond to the behavior of the growing
mode, D+ (a) (there is no growth inside the Horizon during radiation-domination).
96 J. R. Pritchard
10.4.: The processed power spectrum of density perturbations
Matter-Radiation Equality
Density Perturbation
Scale enters
horizon during RD
D!a
D=const
Scale enters
horizon during MD
D!a2
Scale Factor
Figure 10.1: Growth of different scales, entering the horizon during radiation domination
(blue, upper) and matter domination (red, lower).
Cosmology 97
10 — Structure Formation
(t1/2 t−1 )enter ∝ 1/aenter . Hence scales entering during RD experience a relative deficit of
growth by the factor
a2enter
T (k) ∝
a2eq
2
keq
∝ 2 ∝ k −2 k > keq . (10.60)
k
Combining our two cases,
(
1, if k < keq
T 2 (k) = const × (10.61)
k −4 if k > keq .
Traditionally, the time evolution is handled separately, so we set the initial constant to
be one. Thus, an initial spectrum with P (k, ti ) = Ak ns evolves to P (k) = Ak ns on large
scales and P (k) = Ak ns −4 on small scales. (In practice the turnover is not instantaneous,
so the spectrum is smooth near k = keq .) We show how this looks in Figure 10.2 for
ns = 1. (For this specific case, this can also be derived by using the fact that the horizon-
scale perturbation δH , defined in Eq. 10.53 is constant in time, along with the sub-horizon
growth functions of Table 10.4.2.)
3.5
3 P(k) ∝ k
2.5
Log P
1.5
1 P(k) ∝ k-3
0.5
0
0 1 2 3 keq 5 6 7 8
Log k
Figure 10.2: The processed power spectrum of density perturbations, P (k), for an initial
spectrum P (k) ∝ k.
The present-day power spectrum thus depends upon various cosmological parameters.
First, it depends on the value of keq , and hence the ratio of the matter and radiation
densities in the Universe. It obviously depends on the index ns , and hence on some aspects
of inflation. As we will see soon, it also depends on other cosmological parameters through
baryonic features. We see a relatively recent compilation of measurements in Figure 10.3.
98 J. R. Pritchard
f decorrelated power spectra, those for SDSS,
PSCz. Because the power spectra are decor-
fair to do “chi-by-eye” when examining this
similarity in the bumps and wiggles between
Cosmology 99
10 — Structure Formation
perturbations in the pressureless dark matter are able to collapse. It is exactly the
pattern of these baryon-photon sound waves that we see in both the Cosmic Microwave
Background and, at a weaker level, in the matter power spectrum as measured by the
distribution of galaxies. Here, they are known as Baryon Acoustic Oscillations (BAOs).
In fact, the pattern generated by these sound waves is very specific, and they therefore
have a characteristic scale of the maximum distance that a sound wave could have trav-
elled before the baryons and photons decouple at t ' 370, 000 or z ' 1, 100, corresponding
to a scale of roughly 100 Mpc in the Universe today. Indeed, this is the characteristic
scale of spots in the CMB, and more recently we have begun to observe that there are
more galaxies than expected at this separation, a phenomenon known as baryon acoustic
oscillations. Moreover, because this scale is fixed in comoving coordinates by our CMB
observations, it can be used as a standard ruler to measure the cosmological parameters.
The presence of baryons thus modifies our transfer function: in real space, it would
induce a peak at a separation at approximately 100 Mpc; in Fourier space this shows
up as “bumps and wiggles” in T (k) at that scale (roughly speaking, this is because the
Fourier transform of a spike is a sinusoid). These have indeed been observed by the
2DF and Sloan surveys of galaxies on large scales (Figure 10.4), and are beginning to
allow to us to determine cosmological parameters from measurements of the galaxy power
spectrum.
where the P` are Legendre polynomials. Because of linearity, we can in principle ex-
press the CMB power spectrum as a functional of the initial power spectrum of density
perturbations: Z
C` = dkT` (k)P (k, ti ) , (10.68)
where the details of the transfer function depend upon the cosmological parameters.
Just as in the BAO case, the CMB power spectrum lets us determine the cosmological
parameters.
Cosmology 101
10 — Structure Formation
Boltzmann Solvers
In practice, the full description of the CMB and matter power spectrum in the presence
of dark matter, baryons and radiation (as well as neutrinos and possible isocurvature fluc-
tuations as described earlier) requires the solution of the Boltzmann equation, originally
discussed in Chapter 5, but now generalized to allow for spatially-varying perturbations to
the distribution functions and the full treatment of general relativity. This cannot be done
analytically, but instead some powerful numerical codes have been developed in the cos-
mology community. The two most well-known and often-used are CMBFAST (http://
cfa-www.harvard.edu/~mzaldarr/CMBFAST/cmbfast.html) and CAMB (http://camb.
info).
In Figure 10.5 we show the CMB power spectrum for several different sets of cos-
mological parameters as calculated by the CMBFAST program. In Figure 10.6 we show
recent measurements by the WMAP satellite and other experiments, as compiled by the
WMAP team (for more and more recent data, see http://lambda.gsfc.gov).
Figure 10.5: The CMB power spectrum for various sets of cosmological parameters.
102 J. R. Pritchard
10.4.: The processed power spectrum of density perturbations
Figure 10.6: The measured CMB power spectrum, as compiled by the WMAP team.
Cosmology 103
10 — Structure Formation
104 J. R. Pritchard
Errata
Please let me know of any mistakes or typos that you find in the notes so that they can
be corrected in the online version!
Notes
Problems
105