The Standard Model of Cosmology
The Standard Model of Cosmology
The Standard Model of Cosmology
2.1 Introduction
In this chapter we analyse these features in detail, starting with the cosmological
principle in Sect. 2.2. The assumptions of isotropy and homogeneity lead to the
formulation of the FLRW metric, which we introduce in Sect. 2.3. We derive the
dynamic evolution of this metric in Sect. 2.4 by solving the Einstein equation; in
particular, we find that the cosmic expansion is one of the solutions and is favoured by
the measured abundances of the various species. The presence of a cosmic expansion,
in turn, indicates that the primordial Universe was in a very hot and dense state
where thermal equilibrium between the species was established. This prediction is
spectacularly confirmed by the observation of a cosmic microwave background with
a blackbody spectrum, which is the subject of Sect. 2.5. We conclude the chapter by
discussing in Sect. 2.6 some important problems of the hot Big Bang scenario and
one of the possible ways to solve them: the mechanism of cosmic inflation, a phase
of accelerated expansion in the early Universe.
Note that in Sect. 2.6.4 we shall briefly discuss how non-linearities might arise
during inflation that generate non-Gaussian signatures. The work described in this
thesis is ultimately motivated by the quest to measure said non-Gaussianity.
The cosmological principle (CP) states that on sufficiently large scales the Universe
is homogeneous and isotropic. Homogeneity means that different patches of the
Universe have the same average physical properties. In particular, any cosmological
fluid has the same energy density, pressure and temperature everywhere. Isotropy
means that there are no preferred directions in the Universe. Any observer measuring
a cosmological quantity—e.g. the photon flux or a galaxy count—in two different
directions should find the same value.
Homogeneity does not imply isotropy. For example, a Universe filled with a
homogeneous magnetic field is homogeneous but not isotropic. On the other hand,
isotropy about one location does not guarantee homogeneity. The simplest case is
given by an observer at the centre of an isotropic explosion, but there are other
examples of inhomogeneous distributions that project isotropically on the sky of one
observer [25]. However, isotropy about two locations does guarantee homogeneity
and isotropy about all locations (Peacock [62, pp.65–67]).
The cosmological principle is spectacularly violated on small scales. Planets, stars
and galaxies should not exist in a perfectly homogeneous Universe. However, when
zooming out on scales larger than roughly 100 h−1 Mpc, where 1Mpc = 3.086 ×
1022 m = 3.262 × 106 ly is roughly the average distance between two galaxies, the
Universe does become smooth, as we detail in Sect. 2.2.1. This allows us to treat
the dynamics of the cosmological fluids on the largest scales as if the Universe were
perfectly homogeneous and isotropic. In this limit, the physics and the resulting
equation are particularly simple, as discussed in Sect. 2.3.
The cosmological principle also allows us to define a universal time variable, the
cosmic time, defined as the time measured by observers at rest with respect to the
matter in their vicinity. The homogeneity of the Universe ensures that the clocks of
these fundamental observers can be synchronised with respect to the evolution of the
universal homogeneous density. We choose the zero of the cosmic time to coincide
with the Big Bang, which we shall introduce in Sect. 2.4. As a consequence, the
cosmic time is interpreted as the age of the Universe.
2.2 The Cosmological Principle 11
In the 1910s Vesto Slipher had noticed by measuring their light spectra that most of
nearby galaxies—or nebulae, as they were called at the time—were quickly receding
from us [75, 76]. In 1929, Edwin Hubble [41] independently confirmed that galax-
ies where receding and found a correlation between their radial velocity and their
distance from us. This observation is encoded in Hubble’s law, whereby there is a
linear relationship between the radial speed with which a galaxy recedes from Earth
and its distance to it:
v = H0 r . (2.1)
2 T.
3 As a comparison consider that the disk of our Galaxy, the Milky Way, which is an average galaxy,
measures just around 30 kpc.
4 It is sometimes thought that Hubble discovered the expansion of the Universe in his 1929 paper.
This was not the case, as the first connection to Lemaître and Friedmann works was made in 1930
2.3 The Expansion of the Universe 13
Universe picture, the receding galaxies are not thought as projectiles shooting away
through space, but as objects at rest in expanding space. Similarly, the recession
speed is not the speed of something moving through space, but of space itself; it
is not a local phenomenon and this is why it can exceed the speed of light without
changing the causal structure of space-time [36].
The value of H0 cannot be predicted by theoretical means: only observation can pin
it down. Since distance measurements are subject to high uncertainty, it is customary
to parametrize the Hubble constant by means of the pure number h:
km/s
H0 ≡ 100 h (2.2)
Mpc
h
= (2.3)
9.77Gyr
h
= GeV (assuming = 1) (2.4)
4.69 × 1041
h
= (assuming c = 1) . (2.5)
2998 Mpc
In his seminal paper, Hubble estimated h ∼ 5. The most accurate local measure-
ments of h to date employ Cepheid variables and Type Ia supernovae in low-redshift
galaxies, and read
at 68 % confidence level. The Planck CMB satellite obtained a more precise value
[67], but it is an indirect estimate as it assumes a cosmological (ΛCDM) model:
(Footnote 4 continued)
by Arthur Eddington and Willem de Sitter. An account by the American Institute of Physics of
the fascinating story behind the discovery of the expansion of the Universe can be found at the
following URL: http://www.aip.org/history/cosmology/ideas/expanding.htm.
14 2 The Standard Model of Cosmology
The dynamics of the expanding Universe are better understood in terms of observers
who are at rest with the Hubble expansion, the so-called comoving observers. Comov-
ing observers perceive the Universe as isotropic and see objects receding from them
according to Hubble’s law. In this section, we shall employ comoving coordinates
defined as the coordinate system where all comoving observers have constant spatial
coordinates, i.e. are static. Any motion in comoving coordinates has the Hubble part
subtracted so that the only velocities are the peculiar ones.
In differential geometry the distance ds between two infinitesimally nearby space-
time points (x 0 , x 1 , x 2 , x 3 ) and (x 0 + d x 0 , x 1 + d x 1 , x 2 + d x 2 , x 3 + d x 3 ) is called
the line element and is defined as
Here gμν (x) is the metric, a (0,2) tensor which determines how distances are com-
puted in the considered space-time manifold. We shall adopt comoving coordinates
and set x 0 = c t where t is the cosmic time.
The metric that describes a homogeneous and isotropic expanding space-time is
called the Friedmann-Lemaître-Robertson-Walker (FLRW ) metric [32, 46, 70, 82].
In comoving coordinates, it is given by
The cosmic time t, introduced in Sect. 2.2, is defined so that the Universe has the
same density everywhere at each moment in time. The scale factor a(t) parametrises
the uniform expansion of the Universe. We express the spatial part of ds 2 so that, in
comoving and spherical coordinates (ρ, θ, φ), it reads
γi j d x i d x j = dρ2 + Sk (ρ)2 dθ2 + sin2 θ dφ2 . (2.10)
With this choice, the quantity dχ2 ≡ γi j d x i d x j has the meaning of a comoving
distance or coordinate distance. The function Sk (ρ) depends on the spatial curvature
of the Universe, which in these models is uniform and is given by k/a 2 . Even before
discussing its form, it should be noted that for radial trajectories (dφ = dθ = 0) the
comoving distance coincides with the radial comoving coordinate.
2.3 The Expansion of the Universe 15
We distinguish three different geometries for the Universe based on the value of
the curvature constant k:
⎧
⎪ ρ flat geometry (k = 0)
⎪
⎪
⎪
⎨
Sk (ρ) = sin(ρ) spherical geometry (k = +1) (2.11)
⎪
⎪
⎪
⎪
⎩ sinh(ρ) hyperbolic geometry (k = −1) .
For k = 0, the comoving distance is just the usual Euclidean distance: dχ2 =
δi j x i x j . The value of the curvature constant k is a free parameter in the FLRW
models and, as the Hubble constant, has to be determined by experiment. Recent
results from the WMAP [39] and Planck [67] CMB satellites constrain the spatial
curvature to be negligible, thus suggesting that we live in a Universe with a flat
geometry. We shall assume k = 0 for the rest of this work. This allows us to choose
coordinates where ρ and χ are lengths (measured in Mpc) and the scale factor is a
dimensionless quantity such that a(t0 ) = 1 [24].
Now that we have introduced the concept of scale factor, Hubble’s law follows
easily. Given an observer at the origin of a spherical coordinate system, we define
the physical coordinates of an object as r = a(t) x, where x = (x 1 , x 2 , x 3 ) are its
comoving coordinates. The distance r = a(t) χ along a radial path is the physical
distance and can be thought as the distance that would be measured by stretching a
tape measure in a uniformly curved surface [36]. There are two contributions to the
velocity d r/dt:
dr 1 da dx
= r + a . (2.12)
dt a dt dt
We project along the radial direction r̂ in order to obtain an expression for the radial
velocity v = d r/dt · r̂:
1 da dx
v = r + a · r̂ . (2.13)
a dt dt
The term a d x/dt · r̂ is the peculiar velocity of the object. For a comoving object
(d x/dt = 0) we obtain the so-called velocity-distance law:
1 da
v = r. (2.14)
a dt
The above equation has the same form of Hubble’s law in Eq. 2.1. From a direct
comparison, we see that the Hubble constant H0 is just the present-day value of the
Hubble parameter defined as
1 da
H ≡ . (2.15)
a dt
16 2 The Standard Model of Cosmology
Conformal time The FLRW metric can be conveniently expressed using the con-
formal time defined as dτ = dt/a:
ds 2 = a(τ )2 −(c dτ )2 + γi j d x i d x j = a(τ )2 ημν d x μ d x ν , (2.16)
where ημν is the Minkowski metric of special relativity and we have assumed flat
space (k = 0). In the following chapters we shall use τ instead of t as the evolution
variable for the cosmological perturbations, and assume units where c = 1. It should
be noted that, for a radial trajectory, the conformal time is equal to the comoving
distance divided by c.
The cosmological data that we extract from the Universe (temperature and polar-
isation maps, galaxy surveys, lensing maps, etc.) rely on the observation of light,
with the exceptions of neutrinos and, possibly, gravitational radiation. It is therefore
crucial to understand how light is affected by the expansion of the Universe.
All physical lengths are stretched by the expansion of the Universe; the wavelength
of a light wave makes no exception. Light emitted by a comoving source at time t
with wavelength λ will be seen by a comoving observer today with a wavelength λ0
given by
λ0 a(t0 )
= .
λ a(t)
As it travels through the expanding Universe, the light emitted from distant objects
experiences an expansion redshift: its spectrum is uniformly shifted to larger wave-
length and lower energies by an amount depending solely on the time of emission,
regardless of whether the light consists of radio waves or gamma rays.
By adopting the same convention as in spectroscopy, where the fractional wave-
length shift (λ0 − λ)/λ is denoted by the letter z, we write the expansion-redshift
law
a(t0 )
1 + z(t) = . (2.17)
a(t)
If we assume that the laws governing the emission and absorption of light do not
change through cosmic evolution, the expansion redshift of a cosmological source
can be inferred from its electromagnetic spectrum. Thanks to spectroscopic galaxy
2.3 The Expansion of the Universe 17
surveys such as 2dF [18], SDSS-II [87], WiggleZ [23] and BOSS [20], we have
now measured the optical spectra of millions of galaxies and thus determined their
redshift.
In an expanding Universe, the sources with the highest redshift are the ones
farthest away from us. Hence, high-redshift objects have to be more luminous than
low-redshift ones for us to be able to see them. The highest-redshift galaxy that
has been spectroscopically confirmed to date has z = 7.51 [29],5 and a candidate
galaxy with z = 11.9 [27] has been recently reported. In a ΛCDM Universe, the
light from these galaxies was emitted about 13 billion years ago and their distance
is now growing at a rate of many times the speed of light.
In the following, we will sometimes use the redshift as a time variable to para-
metrize the evolution of the Universe. This is correct since z is a monotonically
decreasing function of a which in turn, in an expanding Universe, is a monotonically
increasing function of cosmic time. Note also that from Eq. 2.17 it follows that today
(a(t0 ) = 1) the redshift vanishes: z(t0 ) = 0.
The expansion redshift should not be confused with the Doppler effect. The Doppler
effect produces a shift in the observed wavelength of photons because of the relative
motion between source and observer. The recession velocity does not give rise to
a Doppler shift because it does not describe the motion of objects in space, but
the rate at which distances grow in the expanding Universe. Incidentally, this is
why recession velocities can be larger than the speed of light. What gives rise to
the expansion redshift is the wavelength of photons getting stretched during their
trajectory through expanding space. On the other hand, Doppler redshift is generated
by the peculiar velocities of the galaxies, which cannot exceed the speed of light.
A third type of redshift, the gravitational redshift, arises from the fact that the
photons frequencies change as they travel through an inhomogeneous gravitational
field. For example, we expect the light from a cluster of galaxies to be gravitationally
redshifted, as the gravitational field at the centre of the cluster is different from that
on the surface of Earth.
Expansion redshift, Doppler redshift and gravitational redshift coexist in the spec-
trum of galaxies and, in general, of all astrophysical sources. When determining the
expansion redshift of an object, the non-cosmological Doppler and gravitational
redshifts must be subtracted or accounted for in the error budget. The gravitational
redshift is usually not too much of a concern as it shifts the spectrum by just z ∼ 10−3
[36]. However, in the local Universe, say for z < 0.01, the peculiar velocities give
rise to a Doppler redshift of the same order of the expansion one. This is a man-
ifestation of the breakdown of the cosmological principle on small scales due to
5 Note that a galaxy with a spectroscopic redshift of z = 8.6 had been previously reported in
Ref. [44], but it was later found to be a spurious signal in Ref. [12].
18 2 The Standard Model of Cosmology
gravitational instability. For more distant objects, peculiar velocities become negli-
gible with respect to recession velocities and one can trust the measured redshift to
be due to the expansion of the Universe.
In Sect. 2.3.1 we have introduced the concept of comoving distance χ as the dimen-
sionless distance between two spatial points on the comoving grid. The great advan-
tage of χ is that it is constant in time, since its expression only involves comoving
coordinates. On the other hand, the physical distance, given by r = a(t)χ, is the
tape-measure distance on a grid which is not comoving with the expansion, and
hence increases with time.
But how are these theoretical distances related to the measured redshift of an
object? Since redshift is intrinsically related to light propagation, we need to study
the trajectory of photons from a source to us. This is described by the null geodesics
(ds 2 = 0) along a radial path (dφ = dθ = 0),6 which in the case of the FLRW metric
in Eq. 2.9 yields
c
dχ = dt . (2.18)
a
This result is intuitive: the actual speed of a photon does not vary, but its speed with
respect to expanding coordinates is larger when the Universe is small (a < 1). A
photon that was emitted at a time tems and observed at tobs will have travelled a
comoving distance of
tobs c
χ (tems , tobs ) = dt . (2.19)
tems a(t)
6 Itshould be noted that, given the choice of the spatial metric in Eq. 2.10, the comoving distance
for a radial path is just the radial comoving coordinate.
2.3 The Expansion of the Universe 19
c
dχ = − dz , (2.20)
a0 H (z)
where we have defined the dimensionless parameter E(z) ≡ H (z)/H0 [6]. We shall
refer to the above formula as the distance-redshift law; it is important because it
relates the geometry of the Universe (χ and H ) to the measured redshift. By using
the velocity-distance relation v = H0 r and the identity r (t, t0 ) = a0 χ(t, t0 ), we
obtain the velocity-redshift law
v z dz
= , (2.22)
c 0 E(z)
which is key to convert a redshift to the recession velocity at the time of emission.
The distance-redshift and velocity-redshift laws tell us that, in order to infer the
distances and velocities of an object, we first need to know the expansion history
of the Universe H (z) all the way to when the light was emitted. The reason is that
our cosmological observations are limited to the region of space-time included in
our past light cone. We, as observers, do not have access to a the world map but
only to a single world picture taken now and here [36]. The farthest sources in our
world picture emitted their light at a time where the expansion rate was significantly
different from the current value, H0 . Furthermore, the emitted light travelled for a
long time in an expanding Universe. Hence, the measured redshift is related to the
distance covered by the light by the expansion history between emission time and
observation time. z
If the object is very close, however, the integral 0 dz/E(z) can be Taylor
expanded around z = 0 [6]:
z dz E (0) 2 1 2
z − z + 2E (0) − E (0) z 3 + O(z 4 ) , (2.23)
0 E(z) 2 6
where the prime represents a derivative with respect to z. By keeping only the first
term in the expansion, the distance-redshift and velocity-redshift laws become respec-
tively
c z = H0 r (2.24)
and
v = cz . (2.25)
20 2 The Standard Model of Cosmology
In his famous 1929 paper, Hubble interpreted his velocity measurements as peculiar
velocities rather than recession velocities. He used the Fizeau-Doppler formula to
convert redshifts in velocities, which happens to coincide with the z → 0 limit of
the velocity-redshift law. For this reason, some authors prefer to refer to cz = H0 r
as the Hubble’s law (rather than v = H0 r ) in order to keep clear the distinction
between the Doppler redshift and velocity redshift [36].
The Hubble time t H is defined as the inverse of the Hubble parameter. The current
value of the Hubble time is easily obtained from the definition of H0 in Eq. 2.2:
1
t H0 ≡ = 9.77 h −1 Gyr .
H0
Given constant expansion, i.e. d 2 a/dt 2 = 0, the Hubble time is the time needed by
the Universe to double in size. Equivalently, the solution to:
da
a(t1 ) + Δt = a(t2 ) , (2.26)
dt
for a(t2 ) = 2a(t1 ) is Δt = H −1 (t1 ). If the expansion had been constant after the
Big Bang, the Hubble time would be the age of the Universe; to see it, substitute
a(t1 ) = 0 and a(t2 ) = a in the above equation.
In a more realistic model where the expansion rate varies, the Hubble time does
not correspond anymore to the age of the Universe. It rather sets the time-scale for the
expansion of the Universe: in a time comparable to H −1 the expansion parameter
increases noticeably. In the currently accepted accelerating ΛCDM model, t H0 is
still a good proxy for the current age of the Universe. Using Planck cosmological
parameters [67], one finds t0 = 13.817 ± 0.048 Gyr against t H0 14.6 Gyr.
the horizon is a physical distance, not a comoving one. Its comoving counterpart is
obtained by dividing it by the expansion parameter:
c
.
a(t) H (t)
The above quantity, called the comoving horizon, is not to be confused with the
particle horizon, which we define below and represents the maximum distance a
particle could have travelled since the Big Bang until a certain time t.
The distance travelled by a photon from the Big Bang up to a certain time t is known
as the particle horizon. Its expression in comoving coordinates is obtained from Eq.
2.19 by setting tems = 0 and tobs = t:
t dt
χ(t) ≡ c .
0 a(t)
Since the speed of light is the limit velocity, the particle horizon represents the
maximum comoving distance any particle could have travelled up to time t. Note
that the particle horizon is proportional to the conformal time τ appearing in
Eq. 2.16:
7 Note that this behaviour does not invalidate special relativity since expansion is uniform everywhere
in the Universe and therefore no exchange of information is possible as a result of the super-luminar
velocity.
22 2 The Standard Model of Cosmology
In the following we shall use the conformal time and the comoving particle horizon
interchangeably.
At any moment t in the evolution of the Universe, the particle horizon χ(t) is
the maximum extension of the past light cone for all events in the Universe. In
particular, for an observer on Earth, the present-day particle horizon sets the size of
the observable Universe. Its value depends on the cosmological model adopted; for
a ΛCDM model, it roughly amounts to χ(t0 ) 14,000 Mpc. For the same model,
c t0 4,000 Mpc . There is a subtle difference between the particle horizon χ(t)
and the Hubble horizon c/(a H ): the former is a measure of the past light cone of
an event given the previous expansion history, while the latter sets the extent of its
future light cone based on the instantaneous value of H .
In order to derive the time evolution of the scale parameter a(t) we need to relate
the metric with the energy content of the Universe. This is achieved via the Einstein
equation:
1
Rμν − gμν R = 8 π G Tμν , (2.29)
2
where we have set c = 1 and
• Rμν is the Ricci tensor, defined as the self-contraction of the Riemann tensor. It
can be expressed in terms of the Christoffel symbols or affine connection,
μ g μν ∂gαν ∂gβν ∂gαβ
Γ αβ = + − (2.30)
2 ∂x β ∂x α ∂x ν
as
∂Γ α μν ∂Γ α μα
Rμν = − + Γ α βα Γ β μν − Γ α βν Γ β μα . (2.31)
∂x α ∂x ν
a
Γ 0 i j = δi j a a Γ i 0 j = Γ i j0 = δi j
and , (2.32)
a
a
R00 = −3 and Ri j = δi j 2 a 2 + a a , (2.33)
a 2
a a
R = 6 + , (2.34)
a a
where the primes denote differentiation with respect to cosmic time, a = da/dt.
The left hand side of the Einstein equation is called the Einstein tensor G μν and can
be determined using the above relations:
2
a
G 00 = 3 , G i j = −δi j a 2 + 2 a a G i0 = G 0i = 0 . (2.35)
a
where a = γ, b, ν, c, Λ for photons, baryons, neutrinos, cold dark matter and dark
energy, respectively. The fact that the spatial Einstein tensor is diagonal is a direct
consequence of the isotropy of the FLRW metric. The energy-momentum is forced to
be diagonal too, meaning that the cosmological fluids cannot have peculiar velocities
or anisotropic stresses. Therefore, in the simple FLRW model a fluid is characterised
only by its energy density ρ(t) and its pressure P(t).
We shall assume that the fluids that compose the Universe are barotropic, that is,
their pressure is given as an explicit function of their energy density. The relation
between P and ρ is called the equation of state of the fluid; we parametrise it via the
barotropic parameter w as
P = w(ρ) ρ . (2.37)
As we shall soon see, knowing the equation of state w(ρ) of the various species is
needed to derive the expansion history of the Universe. Relativistic species (R), such
as the photons, the neutrinos and the massive species while still relativistic, have a
constant equation of state: w R = 13 . Non-relativistic species (M), such as the baryons
and cold dark matter after decoupling, instead, have no pressure: wM = 0. Note that,
24 2 The Standard Model of Cosmology
The time-time component of the Einstein equations is called the Friedmann equation,
8π G k
H2 = ρ − 2 , (2.39)
3 a
where H = a /a is the Hubble parameter and ρ = ρa is the total energy density
of the Universe. We have included the curvature contribution, k, to highlight the fact
that in a flat universe (k = 0) the total density always equals the critical density ρcrit ,
defined as
3 H2
ρcrit ≡ .
8π G
The critical density depends on time; its present-day value can be easily computed
in terms of the Hubble constant:
kg
ρcrit (t0 ) = 1.878 h2 × 10−26 (2.40)
m3
M
= 2.775 h−1 × 1011 3 (2.41)
h −1 Mpc
GeV
= 10.54 h2 (assuming c = 1) . (2.42)
m3
This is an astonishingly small number: with a density of 1.27 kg/m3 , air is around
1026 times denser than the critical density. However, since 1011 –1012 solar masses
is close to the mass of a typical galaxy and 1 Mpc is the order of magnitude of the
typical galaxy separation, the Universe cannot be too distant from the critical density.
The density of the species normalised to the critical density of the Universe is
called the density parameter:
ρa (t)
Ωa (t) ≡ . (2.43)
ρcrit (t)
Using the information on the equations of state of the various species (Sect. 2.4.3),
the Friedmann equation can be recast in terms of the present-day value of the density
parameters, Ωa0 ≡ Ωa (t0 ), as
2.4 The Background Evolution 25
ΩM0 ΩR0 Ωk0
H 2 = H02 + + 2 + ΩΛ0 , (2.44)
a3 a4 a
ρM (t0 ) ρR (t0 ) k Λ
ΩM0 = , ΩR0 = , Ωk0 = − , ΩΛ0 = .
ρcrit (t0 ) ρcrit (t0 ) a02 H02 3 H02
(2.45)
(In this thesis, cosmological quantities indexed by a ‘0’ are evaluated today, X 0 ≡
X (t0 )).
a 4π G
= − (ρ + 3 P) , (2.46)
a 3
where P = Pa is the combined pressure of all the species. The acceleration
equation holds also in a curved Universe, where k = 0.
The pressure and the density appear in the acceleration equation on equal grounds:
they both contribute to increasing the gravitational attraction and thus decelerate the
cosmic expansion. This might seem counter intuitive, as we are used to thinking of
pressure as something that powers expansive processes such as explosions. This is
indeed true if a force is supplied by means of a gradient in the pressure field; however,
in a homogeneous Universe, P is the same everywhere and no pressure forces are
possible.
The evolution of the matter species is determined by the conservation of the energy
and momentum,
T μ ν;μ = ∂μ T μ ν + Γ μ αμ T α ν − Γ α νμ T μ α = 0 . (2.47)
ρ + 3 H (ρ + P) = 0 , (2.48)
26 2 The Standard Model of Cosmology
ρ + 3 H ρ (w + 1) = 0 . (2.49)
The continuity equation applies separately to each species as, for the epochs of
interest, their particle number is conserved and their energy exchange is negligible.
Then, for a fluid ‘a’ with a constant equation of state, P = wρ, the continuity
equation can be solved to yield
ρa ∝ a − 3 (1 + wa ) . (2.50)
ρR ∝ a −4 , ρM ∝ a −3 , ρΛ = constant . (2.51)
In the more general case of a time-dependent equation of state, w = w(a), one has
to solve the following integral:
a dã
ρ ∝ exp −3 1 + w(ã) . (2.52)
0 ã
The expansion history of a universe filled by a single species with constant equation
of state can be inferred analytically. This is achieved by inserting the general equation
of state (Eq. 2.50) into the Friedmann equation (Eq. 2.39) and solving for a(t). If the
curvature k is neglected, we have that [24]
which yields a time integral that is easily solved for a(t) once the cosmological
parameters are specified. These have been measured to high accuracy. For the Hubble
2.4 The Background Evolution 27
where we have used the fact that the massless neutrino density is roughly equal to
0.68 Ωγ because they are fermions rather than bosons and are at a lower temperature.
Finally, we assume a flat Universe (Ωk = 0) so that the density of dark energy can
be determined as
In Fig. 2.1 we show the evolution of the scale factor obtained for the above para-
meters. Depending on the species that is the most abundant, we identify three epochs
in the cosmic history: the radiation dominated era (a ∝ τ ), the matter domination
Fig. 2.1 Cosmic history of the Universe. The blue curve is the scale factor as a function of conformal
time, obtained by solving the Friedmann equation in Eq. 2.54. Today corresponds to a = 1 and τ =
14,200 Mpc. The three black dot-dashed curves are the density parameters of radiation (Ω R ), cold
matter (Ω M ) and dark energy considered as a cosmological constant fluid (ΩΛ ). The intersections
between the three Ω’s naturally split the cosmic history in three epochs: the radiation domination
era (a ∝ τ ), the matter domination era (a ∝ τ 2 ) and the dark energy domination era (a ∝ 1/τ )
28 2 The Standard Model of Cosmology
era (a ∝ τ 2 ) and the dark-energy dominated era (a ∝ 1/τ ). The transitions between
the three eras take place at
Ω R0 Ω M0
aeq = = 2.96 × 10−4 and aΛ = = 0.44 , (2.59)
Ω M0 ΩΛ0
a 4π G
= − ρ (3 w + 1) , (2.60)
a 3
we see that in the early Universe when radiation dominates (w = 1/3 > 0), the
second derivative of a(t) is negative; that is, a(t) is a concave curve. Thus, we expect
the scale factor of the Universe to cross the a = 0 line in a finite amount of time;
the moment when this happens is called the Big Bang.8 The Big Bang represents a
singularity in the coordinates (the spatial metric vanishes for a = 0), in the Ricci
scalar (Eq. 2.34) and in the density (ρR ∝ a −4 ).
Soon after the Big Bang, the particle density is so high that the species interact at a
rate much higher than the expansion rate, with all kinds of particle-antiparticle pairs
being created and annihilated. As a result of these continuous collisions, particles of
different species are in thermal equilibrium, i.e. they can be considered to be part of
a single cosmic plasma with a common temperature and average kinetic energy.
Photons in thermal equilibrium obey a blackbody spectrum, which is characterised
by a simple relation between the energy density ργ and the ambient temperature T ,
ργ = α T 4 , (2.61)
where the proportionality constant is the Stefen-Boltzmann constant times 4/c, that
is, α = π 2 kB4 /(3 c3 ). Since the energy density of radiation scales with a −4 , it follows
that the temperature of the cosmic plasma scales as a −1 :
2.725 K
T = = (z + 1) 2.35 × 10−4 eV , (2.62)
a
where we have used the current CMB temperature as normalisation and, in the second
equality, we have assumed units where the Boltzmann constant kB = 11,605−1 eV/K
8 The name was invented during a radio interview by Fred Hoyle, the main supporter of a steady
state Universe, as a mockery of the idea of an expanding Universe. Refer to the following URL for
the transcript: http://www.joh.cam.ac.uk/library/special_collections/hoyle/exhibition/radio/.
2.5 The Cosmic Microwave Background 29
is equal to one. To give an idea of the scales involved, we can use the fact that a ∝ t 1/2
in the radiation dominated era to write
1s 1s
T 1.5 × 10 K 10
1.3 MeV , (2.63)
t t
Thus, one second after the Big Bang, the average photon has an energy of ∼1 MeV
while, after 50,000 years, its energy has dropped to 1 eV.
In an expanding Universe, however, thermal equilibrium does not last forever. The
particles of a given species interact with a rate proportional to their number density,
which decays as a −3 . The expansion rate H , on the other hand, never decays faster
than a −3/2 (Eq. 2.53), meaning that, eventually, it will exceed the interaction rate.
As a result, the thermal equilibrium cannot be maintained anymore and the particle
species is said to have decoupled from the cosmic plasma. As we shall see in the next
sections, the photons decouple at a redshift of z 1100, soon after matter-radiation
equality. Then, why do we speak of “temperature of the photons”, if they are not
in thermal equilibrium? The answer is simple: the cosmic expansion preserves the
blackbody spectrum of the photon fluid even when it is out of thermal equilibrium.
Due to its E/T dependence, the distribution function is frozen as it redshifts into
a similar distribution with a lower temperature proportional to 1/a (we will come
back to this point in Sect. 4.3.1). Thus, after decoupling, the photon fluid possesses
an effective temperature rather than a thermodynamical one.
The presence of this blackbody, isotropic background radiation of cosmic origin is
a definite prediction of the Big Bang model. The first measurement that was directly
linked [21] to the cosmic background radiation was made serendipitously in 1963 by
Penzias and Wilson [65], who measured an isotropic excess temperature of around
3.5 K. Since then, many experiments were performed to measure the present-day
CMB spectrum over different wavelengths. The most accurate measurement of the
CMB spectrum was made by the FIRAS experiment, launched in 1989 on board of the
NASA Cosmic Background Explorer (COBE). The spectrum measured by FIRAS
[30, 54] is blackbody to high accuracy and is shown in Fig. 2.2. The blackbody
form of the CMB spectrum has been confirmed by several other experiments for
wavelengths outside the millimetre range, as shown in Fig. 2.3. The measured CMB
temperature, T0 = 2.725 ± 0.001 K [30], implies that the average CMB photon has
the following properties:
After the temperature of the cosmic plasma has dropped below the electron mass,
T
511 keV, the only process that maintains the photons in thermal equilibrium are
the rapid collisions with the free electrons. In general, the scattering of a photon by
30 2 The Standard Model of Cosmology
Fig. 2.2 The cosmic microwave background spectrum as measured by FIRAS. The error bars have
been multiplied by 400 to make them visible; the line represents the best-fit blackbody spectrum
at T = 2.725 K. Source Data from FIRAS [30], image courtesy of Edward L. Wright from the
website http://www.astro.ucla.edu/~wright/cosmo_01.htm
Fig. 2.3 The CMB blackbody spectrum as confirmed by measurements over a broad range of
wavelengths. Credit: Fig. 19.1 of Ref. [61], reproduced with permission of Springer Publishing
(http://pdg.lbl.gov/1998/contents_large_sports.html); coloured additions courtesy of Karl-Heinz
Kampert (http://astro.uni-wuppertal.de/~kampert/Cosmology-WS0607.html)
2.5 The Cosmic Microwave Background 31
Δλ = λc (1 − cos θ) , (2.65)
which means that the fractional change in the photon’s energy is negligible as long
as its energy is much smaller than the target’s mass. The condition definitely applies
to our context, where we consider temperatures of the order of the eV and the target
particles are electrons with m e c2 = 511 keV.9 In this limit, the process is elastic and
is called Thomson Scattering.
The total cross-section for the Thomson scattering is given by [22]
2
8π 2 2 8π α
σT = α λc = (2.67)
3 3 mc
= 6.652 × 10−29 m2 (2.68)
−17 −2
= 4.328 × 10 eV (assuming h = c = 1 ) , (2.69)
where α 1/137 is the fine structure constant and in the last equalities we have used
the electron mass m e c2 = 511 keV. It is important to note that the cross section is
inversely proportional to the squared mass of the target particle. Therefore, provided
that protons and electrons have the same number density, photon-electron collisions
(m e c2 = 511 keV) are several million times more likely that photon-proton collisions
(m p c2 = 938 GeV). For this reason, we shall ignore the latter and focus on the former.
Here we introduce the interaction rate κ̇ and the optical depth κ that will be useful
in the following chapters to derive and numerically solve the Boltzmann equation.
The cross-section σ associated with a scattering process is defined so that
d N = n σ dx (2.70)
9 Note that, in the context of the cosmological perturbations, even this tiny energy transfer has to be
is the average number of scatterings the incident particle undergoes when covering
a distance of d x in a material with a density n of scattering targets. Since d N /d x
is the average number of scatterings per unit of length, its inverse is the mean free
path:
1
λ = , (2.71)
nσ
i.e. the average distance a particle covers between two consecutive scatterings. If the
velocity d x/dt of the incident particle is known, then it is straightforward to obtain
the interaction rate d N /dt, that is the average number of scatterings per unit of time.
For a photon,
dN
= nσc . (2.72)
dt
The inverse of the interaction rate is the average time elapsed between two consecutive
scatterings; we shall call this quantity mean free time. For a photon it is given by:
1
tγ = . (2.73)
nσc
In the context of the cosmic microwave background, the optical depth or optical
depth, κ, is the average number of Thomson scatterings a photon undergoes from
the time t up to now,
t0
κ(t) = dt n e σT c . (2.74)
t
The optical depth is a monotonically decreasing function of time; its time derivative
is just the interaction rate with a negative sign
dκ
= −n e σT c . (2.75)
dt
In terms of conformal time, dτ = dt/a, the interaction rate reads
dκ
κ̇ = = −a n e σT c . (2.76)
dτ
The frequent Thomson scatterings between the photons and the electrons before
recombination keep the two fluids in thermal equilibrium. Together with the protons,
2.5 The Cosmic Microwave Background 33
which are tightly coupled with the electrons via Coulomb scattering, the three species
form a unique fluid with a common temperature.
The photons are maintained in thermal equilibrium as long as their interaction
rate with the electrons, n e σT c, exceeds the cosmic expansion rate, H . If we assume
that the electrons remain free throughout cosmic evolution, such decoupling happens
only at a redshift of z ∼ 40 [22]. The electrons, however, do not stay free as it is
energetically favourable for them to combine with the free protons to form hydrogen
atoms via the reaction
In the early Universe, the energy and the density of photons are so high that the
hydrogen atoms thus formed are rapidly disrupted via the inverse reaction; thus,
most of the electrons are free and the abundance of neutral hydrogen is very low. As
the Universe expands and cools, however, more and more atoms are able to form and
endure in a process that is called recombination.
During recombination, the number density of free electrons quickly drops and
so does the rate of photon scatterings, |dκ/dt| = n e σT c. When the interaction rate
is surpassed by the expansion rate, the photon fluid goes out of equilibrium and
decouples from the electron fluid. As a result, the photons can stream freely in a
now transparent Universe. This process is called decoupling. As we shall see below,
decoupling happens during recombination.
Recombination is a complicated process that involves non-equilibrium physics
and is usually treated using the Boltzmann formalism. In principle, to obtain the
ionisation history of the Universe requires solving a system with 300+ differential
equations, one per energy level of the hydrogen atom [73]. In practice, however, one
can model the hydrogen atom as having effectively three energy levels: ground state,
first excited state and continuum [63] (see also Sect. 5.3.4). Numerical codes such as
RECFAST [73] start from this 3-level approximation to compute the ionisation history
of the Universe in less than a second with sub-percent accuracy over a wide range of
redshifts. The code HyRec [4] implements an even more accurate numerical treatment
of recombination where four energy levels are considered that is mathematically
equivalent to the multi-level approach [3].
However, it is still possible to make general statements about recombination and
decoupling without resorting to a numerical computation, and we shall do so in
the following two subsections. One of the major simplifications that we shall adopt
is to assume that all the protons are in hydrogen nuclei, thus ignoring the ∼25 %
contribution in mass that is expected from the helium nuclei. Since about 1 proton
out of every 8 is in a Helium nucleus, this results in an error of roughly 10 %.
34 2 The Standard Model of Cosmology
2.5.2.1 Recombination
where n e , n p and n H are respectively the number densities of free electrons, free
protons and neutral hydrogen atoms; note that, since the Universe is globally neutral,
n e = n p . If we neglect the small number of electrons and protons in Helium nuclei,
the denominator is equal to the number density of baryons: n e + n H n b .
Before recombination begins, the reaction e + p ←→ H + γ is in equilibrium
and we use the Saha ionisation equation [22, 24] to describe it:
3/2
xe2 1 me T
= e−/T . (2.79)
1 − xe ne + n H 2π
If we approximate n e + n H n b and multiply and divide the right hand side by the
blackbody density of the photons, n γ = 2/π 2 T 3 ζ(3), where ζ(3) 1.2021, we
obtain
Because of the steep slope of the xe curve, these values are not particularly sensitive
to the choice of xe (Trec ). It should be noted that Trec is considerably smaller than
the energy needed to ionise an hydrogen atom. The reason is that the large value
of n γ /n b pushes xe to unity and significantly delays recombination; the photons
are so abundant that, even at sub-eV energies, there are still enough of them in the
high-energy tail of the Planck distribution to keep the Universe ionised [24].
The Saha equation is meant to be accurate only when recombination happens in
quasi-equilibrium. In Fig. 2.4, we show the Saha solution together with the “exact”
ionisation history as obtained from solving the Boltzmann equation. As expected,
the Saha approximation is accurate in determining the redshift when recombination
starts but it fails at lower redshifts when the system goes out of equilibrium. It should
be noted that the xe curve flattens at low redshift, as if recombination at some point
2.5 The Cosmic Microwave Background 35
Fig. 2.4 Ionisation history of recombination. The free electron fraction is plotted against redshift
and temperature. Recombination starts when xe begins to drop and is a quick process. The Saha
approximation (Eq. 2.80) correctly describes the beginning of recombination, but fails when the
average energy of the photons becomes too small to maintain the e + p ↔ H + γ reaction in
equilibrium. Note that the exact solution does not drop to zero but, due to the reaction “freezing”
when σT xe n b c
H , it asymptotes to xe 10−3 . Source Dodelson [22, p. 72], reproduced with
permission from Elsevier Books
had become ineffective in binding electrons and protons. This is indeed what happens
after the recombination rate drops below the expansion rate, so that recombination
“freezes” and the ionisation fraction remains constant.
2.5.2.2 Decoupling
Two particle species decouple from each other when their interaction rate drops
below the cosmic expansion rate. Roughly speaking, if a photon scatters an electron
less than once in an expansion time, equilibrium between the two species cannot
be maintained. As we mentioned above, all the species are doomed to decouple at
some point due to the expansion rate decreasing slower than any interaction rate.
For the photons, the process of recombination anticipates this moment by suddenly
removing most of the free electrons from the Universe.
We estimate the redshift of photon decoupling by equating the rate of photon
scatterings with the cosmic expansion rate:
Provided that we neglect the helium nuclei, we can express the fraction of free
electrons as
Ωb0 ρcrit
n e = xe n b = xe (1 + z)3 .
mp
36 2 The Standard Model of Cosmology
where we have neglected the cosmological constant and the curvature because they
were insignificant at the high redshifts considered. By enforcing the condition in Eq.
2.82 we obtain
−1/2 1/2
1 + z dec m p H 0 Ω
xe (1 + z dec )3/2 1 + = M0
. (2.84)
1 + z eq c ρcrit σT Ωb0
Inserting the cosmological parameters considered in Sect. 2.4.4, the term in the right
hand side evaluates to 236 and z eq 3380. The ionisation fraction xe needs to be
computed numerically (Saha’s equation is of no use when xe is small) and we do so
by using RECFAST [73]. This results in the values z dec 900 and xe (z dec ) 10−2 ,
which imply that photon decoupling takes place during recombination (recombina-
tion ends when the ionisation fraction reaches the freeze-out value of xe 10−3 , see
Fig. 2.4). It is interesting to note that if recombination did not happen the photons
would have decoupled only at z 40; this can be seen by setting xe = 1 in the above
equation.
In Sect. 5.5 (and in SONG) we shall use a more sophisticated method to determine
the time of photon decoupling, making use of the visibility function, the probability
that a photon last scattered at a given redshift. In particular, we shall see that the
visibility function peaks at z dec 1100, a redshift slightly higher than what we
have inferred by enforcing n e σT c = H . For a standard ΛCDM model, a redshift of
z dec 1100 correponds to
The standard hot Big Bang model introduced in the previous sections succesfully
accounts for the observed expansion of the Universe (Sect. 2.4.4), for the blackbody
spectrum of the cosmic microwave background (Sect. 2.5) and for the abundances of
the light nuclei created via nuclesynthesis (see, for example Refs. [22] and [24]). The
model, however, is unable to answer several important observational and theoretical
questions that we list below.
• The Big Bang singularity The most obvious issue is the presence of a a sin-
gularity in the finite past, the Big Bang (Sect. 2.4), when the curvature and the
density of the Universe are divergent.
• The Horizon problem Any sign of correlations between regions of the Universe
separated by a distance larger than the particle horizon cannot be explained by
the standard model (Sect. 2.3.5). This is, however, what we observe: the cosmic
microwave background has the same temperature with a precision of a part over
105 regardless of the direction of observation. The particle horizon at decoupling
was 80 times smaller than the current value (Sect. 2.5.2.2), meaning that we would
expect to observe fluctuations of order unity in the temperature of the CMB sky on
angular scales of about 1◦ . The fact that we do not observe such fluctuations poses
a causality problem that is referred to as the horizon problem: how can regions of
the Universe be so similar if they did not have enough time to interact?
• The Flatness problem The Friedmann and acceleration equations (Eqs. 2.44
and 2.46) can be combined to obtain an evolution equation for the total density
parameter Ω(t) ≡ ρ/ρcrit = 1 − k/(a 2 H 2 ):
d
Ω(t) − 1 = Ω(t) − 1 Ω(t) (1 + 3w) . (2.86)
dt
This equation shows that, for a Universe with an equation of state of w > −1/3,
such as in a mixture of matter and radiation, the solution Ω(t) = 1 is dynamically
unstable; in fact, the sign of the derivative is positive for Ω(t) > 1 and negative
for Ω(t) < 1 so that Ω(t) will always evolve away from unity. This means
that, for the Universe to be close to the critical density today as observations
suggest, it had to be much more so in the past. For example, for a current value
of 0.1 < Ω0 < 2, it can be shown [24] that |Ω − 1| ≤10−15 at nucleosynthesis
(z 109 ) and |Ω−1| ≤ 10−60 at the Planck time (t P = G/c5 5.4×10−44 s).
The smallness of these values poses a fine-tuning issue that is called the flatness
problem: how can the Universe be still so close to the critical density?
• The structure problem We observe tiny anisotropies in the CMB with an ampli-
tude of ΔT /T ≈ 105 and, more evidently, the observed Universe is highly inho-
mogeneous with a strongly clustered distribution of galaxies on small scales. By
which mechanism was this structure formed?
These shortcomings of the hot Big Bang model are all connected to the initial
conditions of the Universe. In this section we shall see that, apart from the Big Bang
38 2 The Standard Model of Cosmology
The mechanism of cosmic inflation [2, 33, 50, 77] consists of postulating the exis-
tence of a period in which the Universe was much smaller than what one would infer
based on the standard Big Bang model. In this period, the same regions of the Uni-
verse that we see today as separate and independent, were actually in causal contact.
In order to link this “small universe” with the size of the universe today, one needs
to postulate a phase in between where the Universe has expanded much quicker than
the normal rate; hence the name cosmic inflation. In Fig. 2.5 we explain this process
in terms of a conformal diagram of cosmic inflation.
Cosmic inflation solves the horizon problem by connecting regions that, in a
standard Big Bang model, would be causally disconnected. For this to happen, the
comoving Hubble radius, which we defined in Sect. 2.3.5 to be c/(a H ), at the
beginning of inflation had to be larger than the largest scale observable today, that
is the current comoving Hubble radius. Since after inflation the horizon grows with
time (Sect. 2.4.4), it follows that during inflation it has to decrease; the expansion
during inflation must therefore satisfy
d 1 d 2a
< 0 ⇒ > 0, (2.87)
dt aH dt 2
that is, the expansion had to be accelerated. It is important to remark that it is not the
accelerated expansion that solves the horizon problem: the causal connection (i.e. the
2.6 Cosmic Inflation 39
Fig. 2.5 Conformal diagram of inflation. The y-axis is conformal time, while the x-axis is distance.
Our vantage point is today (τ0 ), on the x = 0 vertical line. The standard Big Bang model predicts that
the dynamical evolution of the Universe started at τ = 0 (green horizontal line). In this picture, the
past light cones of two distant CMB patches (small orange triangles) do not intersect, because the
particle horizon at the time where the CMB is formed (horizontal line at τrec ) is much smaller than
τ0 . Therefore, we expect order-unity differences in the CMB temperature on large scales. However,
we observe the CMB today to be almost perfectly isotropic on all scales; this is the horizon problem.
In the inflationary scenario, the horizon problem is solved by postulating the existence of a period
where the two CMB patches were in casual contact (big orange triangle). This is achieved by
extending the time axis below τ = 0 in order to allow the past-light cones of the two CMB patches
to intersect. A period of accelerated expansion, cosmic inflation, is needed in order to bridge the
gap between the “small Universe” where the casual contact was established, and the large value
of today’s particle horizon. In this context, τ = 0 is not a singularity but an apparent Big Bang,
as it marks the end of inflation and the decay of the inflaton (Sect. 2.6.2) into a thermal mix of
elementary particles. The actual Big Bang singularity sits at τ → −∞. Source Courtesy of Daniel
Baumann, from Fig. 9 of Baumann [10]
Universe becoming uniform) is established before inflation and what inflation does
is to put those regions out of reach again, because this is how we see them today.
The accelerated expansion, however, does solve the flatness problem, because
it washes out any curvature, stretching the geometry of the Universe so much that
40 2 The Standard Model of Cosmology
it becomes spatially flat [38]. More quantitatively, we see from the acceleration
equation (Eq. 2.46),
a 4π G
= − (ρ + 3 P) , (2.88)
a 3
that the Universe undergoes an accelerated expansion only if ρ + 3 P < 0 or, in
terms of the barotropic parameter, if w < − 13 . If we inspect Eq. 2.86, we realise that
this is the same condition needed to make Ω(t) = 1 an attractor solution; that is, if
cosmic inflation lasted long enough, the flatness problem would be solved without
the need to fine tune the initial curvature. In fact, we can ask the question: how many
times must the Universe double in size during inflation to justify the fact that today’s
Universe is so close to the critical density? The answer comes from the Friedmann
equation for a constant equation of state (Eq. 2.44):
3 |k|
|Ω(t) − 1| = ∝ a1 + 3 w . (2.89)
8 π G a2 ρ
If we assume that during the inflationary phase w = −1, then |Ω(t) − 1| decreases
like a −2 ; to bring |Ω(t) − 1| to today’s value of order unity from ∼10−60 at the
Planck time would require that
aend
N ≡ ln 30 ln(10) 70 , (2.90)
aini
where N is called the number of e-foldings and aini and aend mark the beginning and
the end of inflation, respectively.
Cosmic inflation provides a solution to the structure problem that is rooted in
quantum mechanics; we postpone this discussion until Sect. 2.6.3.
Inflation is a mechanism rather than a theory of the early Universe, a phase of accel-
erated expansion before which the comoving horizon was larger than the largest
scale observable today. We have seen that to realise the accelerated expansion it
is necessary for the matter content of the Universe to have an equation of state of
w < − 13 , which corresponds to a negative pressure, ρ+3 P < 0. Neither cold matter
(w = 0) nor radiation (w = 13 ) are suitable candidates as they have positive pressure;
the cosmological constant (w = −1) can produce an accelerated expansion but is
completely negligible in the early Universe, so it cannot be responsible for inflation.
Let us see how the presence of a scalar field, which we call the inflaton φ, can
trigger the mechanism of cosmic inflation. The scalar field Lagrangian is given by
2.6 Cosmic Inflation 41
1
Lφ = − ∂μ φ ∂ μ φ − V (φ) , (2.91)
2
where V (φ) is the potential for the field, which we assume to be positive. In principle
L should include terms to account for the interactions with the other species, but we
postulate that they are negligible during inflation. The pressure and the energy density
of the inflaton field can be inferred from its energy-momentum tensor:
1
Tμν = ∂μ φ ∂ν φ − gμν ∂α φ ∂ α φ − gμν V (φ) . (2.92)
2
Here we assume that the Universe is homoegenous, so that gμν is the conformal
FLRW metric in Eq. 2.16 and the spatial gradients of φ vanish. It follows that
1 2 1 i 1
ρφ = −T 0 0 = φ + V (φ) and Pφ = T i = φ 2 − V (φ) .
2 3 2
(2.93)
where φ = dφ/dt. The expression for the energy density is reminiscent of that
of a particle moving in a potential V with velocity φ and kinetic energy 21 φ 2 . In
this picture, a field with negative pressure is one with more potential energy than
kinetic. In the limit where the inflaton field is constant (φ = 0), its kinetic energy
vanishes and we have a constant energy density: ρφ = V (φ) = constant. If we
assume that the energy density and pressure of the Universe are dominated by the
inflaton’s contribution, the expansion rate of the Universe is determined by ρφ via
the Friedmann equation Eq. 2.39:
1 da 8 π G ρφ
H = = = constant , (2.94)
a dt 3
where we have introduced the Planck mass m P ≡ (8πG)−1/2 2.4 × 1018 GeV.
The Friedmann and acceleration (Eq. 2.46) equations can be combined to yield the
background evolution of the inflaton,
where the primes denote derivatives with respect to cosmic time t and V,φ = ∂V /∂φ.
42 2 The Standard Model of Cosmology
Fig. 2.6 Example of a slow-roll inflationary potential. As long as the inflaton’s kinetic energy, 21 φ 2 ,
is negligible with respect to its potential energy, V (φ), the Universe expands in an accelerated
fashion; this limit corresponds to the constant part of the potential. When 21 φ 2 V (φ), the
acceleration can no longer be sustained and inflation ends. When the inflaton reaches the minimum
of the potential, reheating occurs and the energy density of the inflaton is converted into a thermal
mix of elementary particles. Source Courtesy of Daniel Baumann, from Fig. 10 of Baumann [10]
We have just proved that a scalar field can drive inflation as long as it does not evolve
significantly, φ 2
V (φ). The issue now is to determine the potential V (φ) that
keeps φ nearly constant for the number of e-foldings necessary to solve the horizon
and flatness problems. Most models of inflation satisfy the slow-roll condition [2,
50], whereby the inflaton stays nearly constant by slowly rolling down a potential
that is almost flat. We show an example of a slow-roll potential in Fig. 2.6. Because
inflation cannot last forever, the potential needs to have a minimum; as time goes
on, the inflaton approaches this minimum and, due to the increased slope of the
potential, it starts to evolve faster. Inflation comes to an end when the kinetic energy
1 2
2 φ grows to be of the order of the potential V (φ). When the inflaton eventually
reaches the minimum of the potential, the coupling with the other fields becomes
significant so that it decays into a thermal mix of elementary particles [24], leading
to a radiation dominated universe in a process called reheating. In practice, we can
think of the reheating process after inflation as the moment when the hot Big Bang
occurs, in which matter and radiation as we know them start to be created.
Many different potentials can be devised that satisfy the slow-roll condition. It
is customary to parametrise them with two variables that vanish in the limit where
φ is constant. The first slow-roll parameter η quantifies the variation in the Hubble
factor, and is related to the first derivative of the inflaton potential. It is defined as
2
d 1 H m 2P V,φ
≡ = − 2 ≈ . (2.97)
dt H H 2 V
2.6 Cosmic Inflation 43
√
Whenever the inflaton field is constant, φ = 0, then also H ∝ ρφ is constant (Eq.
2.94) meaning that the parameter vanishes. In fact, the slow-roll condition requires
1, an assumption that implies an approximate time-translation invariance of the
background. On the other hand, in the radiation dominated era = 2; in fact, one can
define the inflationary epoch as < 1. The second slow-roll parameter, η, is directly
related to the second derivative of the potential,10
V,φφ
η ≡ m 2P . (2.99)
V
Again, in the case of a constant field or potential this parameter vanishes. As we shall
see below, the most important predictions of inflation can be recast in terms of the
slow-roll parameters and η.
Cosmic inflation was originally proposed to solve the horizon and flatness prob-
lems [2, 33, 50, 77], but it was soon realised that it also provided a mechanism to
generate primordial density fluctuations [7, 37, 59, 78]. The idea is that the struc-
ture that we observe today, such as the CMB anisotropies and the galaxy distribution,
formed starting from tiny quantum fluctuations set during inflation and later enhanced
throughout cosmic history via gravitational instability. These primordial fluctuations
were generated as microscopic quantum vacuum fluctuations in the inflaton field that,
during inflation, were stretched and imprinted on superhorizon scales by the acceler-
ated expansion. These density fluctuations reentered the horizon after inflation ended
and served as initial conditions for the anisotropy and the growth of structure in the
Universe.
In what follows, we briefly describe the main features of the primordial fluctua-
tions generated during inflation. To do so, we need to use some concepts that will be
formally defined only in the next chapter, like the idea that the primordial fluctuations
generated during inflation are stochastic in nature and, therefore, their magnitude is
determined in terms of their variance (in real space) or their power spectrum (in
Fourier space). We will also use of the concepts of scalar and tensor (Sect. 3.3.1)
perturbations (Sect. 3.4), power spectrum (Sect. 3.7.1) and bispectrum (Sect. 3.7.2).
10 In defining the slow-roll parameters, we are using the notation of the review by Bartolo et al. [8].
Chen [15], on the other hand, denotes the quantity in Eq. 2.99 as η V and uses the symbol η for a
third slow-roll parameter:
η ≡ −2 η V + 4 = . (2.98)
H
.
44 2 The Standard Model of Cosmology
The primordial fluctuations generated during slow-roll inflation are expected to have
nearly the same variance on all spatial scales. The reason is that the slow-roll condition
= −H /H 2
1 results into an approximate time-translation invariance of the
background. Therefore, the primordial fluctuations are produced with approximately
the same background expansion rate regardless of the scale considered. This scale
invariance is usually quantified in terms of the scalar spectral index, n s , defined
to be the slope of the dimensionless power spectrum of the primordial curvature
perturbation,
PR ∝ k n s −1 . (2.100)
ns = 1 − 6 + 2 η . (2.101)
Because the slow-roll parameters and η describe, respectively, the first and second
derivative of the inflaton potential V (φ), measuring n s is equivalent to constraining
the shape of V (φ). The cosmic microwave background is strongly affected by the
tilt of the primordial fluctuations and, as a result, it can be used to constrain n s [67]:
This measurement is in agreement with the slow-roll inflationary models and suggests
that the two slow-roll parameters have a value of O(10−2 ).
Another important observable of inflation is the amplitude As of the primordial
fluctuations, which is defined as
n s −1
k
PR (k) = As , (2.103)
k0
where k0 is the pivot scale. In the slow-roll limit, the amplitude As is connected to
the ratio between the inflaton potential and the slow-roll parameter [68]:
V
As = . (2.104)
24 π 2 m 4P
By measuring the amplitude of the CMB angular spectrum, the Planck team [68]
+0.024
found the value ln(1010 As ) = 3.089−0.027 at 68 % confidence level for a pivot scale
2.6 Cosmic Inflation 45
V 1/4
= 0.027 m P = 6.6 × 1016 GeV . (2.105)
1/4
defines the tensor amplitude At and the tensor spectral index n t , which vanishes
for a scale-invariant spectrum. For a slowly rolling scalar field, they are given by
[22, 68]
2V
At = , and n t = −2 . (2.107)
3 π 2 m 4P
In the slow-roll limit, a consistency relation links the spectral index n t to the ampli-
tudes of the scalar and tensor power spectra:
Pt
r ≡ = −8 n t , (2.108)
PR
where we have defined the tensor-to-scalar ratio r . Since As has already been exper-
imentally determined, measuring the value of r would automatically yield the ampli-
tude of the tensor perturbations At and, through the consistency relation, the tilt n t of
the tensor spectrum. Furthermore, a determination of r would imply also an indirect
detection of the gravitational waves. So far, only upper limits for the tensor-to-scalar
exist; in Fig. 2.7 we show the joint measurement of r and n s produced by the Planck
experiment [68].
2.6.4 Non-Gaussianity
The inflation observables that we have introduced in the previous subsection, the
spectral index n s and the tensor-to-scalar ratio r , are defined with respect to the
power spectrum of the primordial curvature perturbation, PR . The power spectrum,
46 2 The Standard Model of Cosmology
0.25
Planck+WP
Planck+WP+highL
0.20
Tensor-to-Scalar Ratio (r0.002 )
Co
nv Planck+WP+BAO
Co ex Natural Inflation
nca
ve
0.15
V ∝ φ2/3
V ∝φ
V ∝ φ2
0.05
V ∝ φ3
N∗ =50
0.00
Fig. 2.7 Constraints on the spectral tilt and the tensor-to-scalar ratio r from Planck [68]. The
ellipses represent the 68 and 95 % confidence limits on n s and r for various combinations of
datasets (WP WMAP polarisation, BAO Baryon acoustic oscillation, highL high-resolution CMB
data). The theoretical predictions of several inflationary models are also shown. Credit: Fig. 1 on
p.10 of Ref. [68] by the Planck collaboration, A&A, reproduced with permission c ESO
however, is just one of the infinite series of n-point functions that characterise the
primordial field (Sect. 3.4). In the case of a Gaussian random field, these moments
can be expressed as products of PR ; for an arbitrary field, this is not the case: the
higher-order moments contain extra information that eludes the power spectrum and
that, as we shall soon see, is precious to understand the non-linear physics at work
in the early Universe. We shall refer to this extra information as non-Gaussianity,
simply because it is absent for Gaussian perturbations.
In this thesis, we focus on the three-point function of the primordial curvature
perturbation, or primordial bispectrum. The full formalism to characterise the bis-
pectrum and its observability in the cosmic microwave background will be introduced
in Chap. 6. The purpose of this subsection is to explain our motivations for studying
the bispectrum; therefore, for now, we shall keep the technical details to a minimum.
The primordial bispectrum is important for two reasons. First, it is the lowest order
statistic sensitive to whether a perturbation is Gaussian or non-Gaussian. This follows
from the fact that the three-point function of a Gaussian random field with zero mean
vanishes. Secondly, it is directly related to the angular bispectrum of the cosmic
microwave background, which is an observable quantity [42, 43, 85]. Therefore, the
primordial bispectrum as inferred from the CMB has the power of discriminating
models of inflation based on the amount of non-Gaussianity they produce.
The standard slow-roll inflation models that we have described above, where the
accelerated expansion is driven by a non-interacting scalar field, produce a bispec-
trum of the order of the slow-roll parameters [1, 53]; for all practical purposes, this
non-Gaussianity can be considered negligible. This is intuitive as the bispectrum is
inherently related to the non-linearities in the propagation of the field. In the “vanilla”
2.6 Cosmic Inflation 47
models, the inflaton propagates freely along a very flat potential (, |η|
1), so that
any self-interaction term of the inflaton potential and the gravitational coupling must
be very small; consequently, the non-linearities are also suppressed [8].
Measuring a significant bispectrum would therefore rule out the simplest models
of inflation. It should be stressed that these models are otherwise highly successful
in reproducing the required duration of inflation and the observed shape of the power
spectrum. The non-Gaussianity measurement is thus complementary to the usual
inflation observables, n s and r , and it provides extra information on the physics of
the early Universe that is useful to break degeneracies between models that would
otherwise be observationally equivalent.
The constraining power of the primordial bispectrum and its observability promp-
ted particle physicists and cosmologists to join forces and investigate many well-
motivated extensions to the inflationary vanilla model. The multiple-field models,
for example, postulate that two or more fields are present during inflation. These mod-
els are appealing also because, from the point of view of particle physics, it is natural
to have several other fields that contribute to the inflationary dynamics. If the fields
interact, the Lagrangian will include non-linear contributions that ultimately lead to
deviations from pure Gaussian statistics [8, 13, 74]. This is not, however, the only
mechanism to create non-Gaussianity in a multi-field model. In the curvaton scenario
[28, 49, 51, 56, 57], for example, the inflaton field drives the accelerated expansion
as in a single field model, while a subdominant second field, the curvaton, is respon-
sible for generating the curvature perturbations. In this case, the non-Gaussianity is
produced by the non-linear evolution of the curvature perturbation on superhorizon
scales.
Other extensions to the vanilla model include features in the inflaton potential, the
presence of a non-canonical kinetic term, non-linearities in the initial vacuum state or
modifications to the theory of gravity [15]. These features generally translate to non-
Gaussian signatures in the primordial curvature perturbation and, thus, in specific
shapes of the bispectrum. For a review on these models and their observability, refer
to the reviews in Refs. [9, 42, 48, 85].
In summary, the non-Gaussianity of the cosmological perturbations opens a win-
dow on the non-linear physics of the early Universe; the CMB bispectrum is the
observable that allows us to look through this window. The subject of this thesis is
the connection between the primordial non-Gaussianity and the CMB bispectrum.
In the following chapters, we shall answer the questions: how is the measured CMB
bispectrum affected by the non-linear evolution that happens after inflation? Would
this effect significantly bias a measurement of the primordial signal?
The answers can be found in Chap. 6.
48 2 The Standard Model of Cosmology
References
20. Dawson KS, Schlegel DJ et al (2013) The Baryon oscillation spectroscopic survey of SDSS-III.
AJ 145:10. doi:10.1088/0004-6256/145/1/10. arXiv:1208.0022
21. Dicke RH, Peebles PJE, Roll PG, Wilkinson DT (1965) Cosmic black-body radiation. Astro-
phys J 142:414–419. doi:10.1086/148306
22. Dodelson S (2003) Modern cosmology. Academic Press, New York
23. Drinkwater MJ, Jurek RJ, Blake C, Woods D, Pimbblet KA et al (2010) The WiggleZ Dark
Energy Survey: survey design and first data release. MNRAS 401:1429–1452. doi:10.1111/j.
1365-2966.2009.15754.x. arXiv:0911.4246
24. Durrer R (2008) The cosmic microwave background. Cambridge University Press, Cambridge
25. Durrer R, Eckmann JP, Sylos Labini F, Montuori M, Pietronero L (1997) Angular
projections of fractal sets. Europhys Lett 40:491–496. doi:10.1209/epl/i1997-00493-3.
arXiv:astro-ph/9702116
26. Ellis GFR (1975) Cosmology and verifiability. QJRAS 16:245–264
27. Ellis RS, McLure RJ, Dunlop JS, Robertson BE et al (2013) The abundance of star-forming
galaxies in the redshift range 8.5–12: new results from the 2012 Hubble ultra deep field cam-
paign. ApJ 763:L7. doi:10.1088/2041-8205/763/1/L7. arXiv:1211.6804
28. Enqvist K, Sloth MS (2002) Adiabatic CMB perturbations in pre-Big-Bang string cosmology.
Nucl Phys B 626:395–409. doi:10.1016/S0550-3213(02)00043-3. arXiv:hep-ph/0109214
29. Finkelstein SL, Papovich C, Dickinson M, Song M, Tilvi V et al (2013) A galaxy rapidly
forming stars 700 million years after the Big Bang at redshift 7.51. Nature 502:524–527.
doi:10.1038/nature12657. arXiv:1310.6031
30. Fixsen DJ, Cheng ES, Gales JM, Mather JC, Shafer RA, Wright EL (1996) The cosmic
microwave background spectrum from the Full COBE FIRAS data set. Astrophys J 473:576.
doi:10.1086/178173. arXiv:astro-ph/9605054
31. Freedman WL, Madore BF, Scowcroft V, Burns C, Monson A, Persson SE, Seibert M, Rigby
J (2012) Carnegie Hubble program: a mid-infrared calibration of the Hubble constant. ApJ
758:24. doi:10.1088/0004-637X/758/1/24. arXiv:1208.3281
32. Friedmann A (1922) Über die Krümmung des Raumes. Zeitschrift fur Physik 10:377–386.
doi:10.1007/BF01332580
33. Guth AH (1981) Inflationary universe: a possible solution to the horizon and flatness problems.
Phys Rev D 23(2):347–356. doi:10.1103/PhysRevD.23.347
34. Guzzo L (1997) Is the universe homogeneous? (On large scales). New Astron 2:517–532.
doi:10.1016/S1384-1076(97)00037-7. arXiv:astro-ph/9711206
35. Hamilton JC (2013) What have we learned from observational cosmology? In: Proceedings of
“Philosophical Aspects of Modern Cosmology” held in Granada, Spain, 2011. arXiv:1304.4446
36. Harrison ER (2000) Cosmology. The science of the universe. Cambridge University Press,
Cambridge
37. Hawking SW (1982) The development of irregularities in a single bubble inflationary universe.
Phys Lett B 115:295–297. doi:10.1016/0370-2693(82)90373-2
38. Hawley JF, Holcomb KA (2005) Foundations of modern cosmology, 2nd edn. Oxford Uni-
verxity Press, Oxford
39. Hinshaw G, Larson D, Komatsu E, Spergel DN, Bennett CL, Dunkley J, Nolta MR, Halpern
M, Hill RS, Odegard N, Page L, Smith KM, Weiland JL, Gold B, Jarosik N, Kogut A, Limon
M, Meyer SS, Tucker GS, Wollack E, Wright EL (2013) Nine-year Wilkinson Microwave
Anisotropy Probe (WMAP) observations: cosmological parameter results. ApJS 208:19.
doi:10.1088/0067-0049/208/2/19. arXiv:1212.5226
40. Hogg DW, Eisenstein DJ, Blanton MR, Bahcall NA, Brinkmann J, Gunn JE, Schneider DP
(2005) Cosmic homogeneity demonstrated with luminous red galaxies. Astrophys J 624:54–58.
doi:10.1086/429084. arXiv:astro-ph/0411197
41. Hubble E (1929) A relation between distance and radial velocity among extra-galactic nebulae.
Proc Natl Acad Sci 15:168–173. doi:10.1073/pnas.15.3.168
42. Komatsu E (2010) Hunting for primordial non-Gaussianity in the cosmic microwave
background. Class Quantum Gravity 27(12):124,010. doi:10.1088/0264-9381/27/12/124010.
arXiv:1003.6097
50 2 The Standard Model of Cosmology
65. Penzias AA, Wilson RW (1965) A measurement of excess antenna temperature at 4080 Mc/s.
Astrophys J 142:419–421. doi:10.1086/148307
66. Pietronero L (1987) The fractal structure of the universe: correlations of galaxies and clus-
ters and the average mass density. Phys A Stat Mech Appl 144:257–284. doi:10.1016/0378-
4371(87)90191-9
67. Planck Collaboration (2014a) Planck 2013 results. XVI. Cosmological parameters. A&A
571:A16. doi:10.1051/0004-6361/201321591. arXiv:1303.5076
68. Planck Collaboration (2014b) Planck 2013 results. XXII. Constraints on inflation. A&A
571:A22. doi:10.1051/0004-6361/201321569. arXiv:1303.5082
69. Riess AG, Macri L, Casertano S, Lampeit H, Ferguson HC, Filippenko AV, Jha SW, Li W,
Chornock R, Silverman JM (2011) A 3% solution: determination of the Hubble constant with
the Hubble space telescope and wide field camera 3. ApJ 732:129. doi:10.1088/0004-637X/
732/2/129
70. Robertson HP (1935) Kinematics and world-structure. ApJ 82:284. doi:10.1086/143681
71. Scharf CA, Jahoda K, Treyer M, Lahav O, Boldt E, Piran T (2000) The 2–10 keV X-ray
background dipole and its cosmological implications. Astrophys J 544:49–62. doi:10.1086/
317174. arXiv:astro-ph/9908187
72. Scrimgeour MI, Davis T, Blake C, James JB, Poole GB et al (2012) The WiggleZ Dark Energy
Survey: the transition to large-scale cosmic homogeneity. MNRAS 425:116–134. doi:10.1111/
j.1365-2966.2012.21402.x. arXiv:1205.6812
73. Seager S, Sasselov DD, Scott D (1999) A new calculation of the recombination epoch. Astro-
phys J Lett 523:L1–L5. doi:10.1086/312250. arXiv:astro-ph/9909275
74. Seery D, Lidsey JE (2005) Primordial non-Gaussianities from multiple-field inflation. J Cosmol
Astropart Phys 9:011. doi:10.1088/1475-7516/2005/09/011. arXiv:astro-ph/0506056
75. Slipher VM (1913) The radial velocity of the Andromeda Nebula. Lowell Obs Bull 2:56–57
76. Slipher VM (1915) Spectrographic observations of Nebulae. Popul Astron 23:21–24
77. Starobinsky AA (1980) A new type of isotropic cosmological models without singularity. Phys
Lett B 91:99–102. doi:10.1016/0370-2693(80)90670-X
78. Starobinsky AA (1982) Dynamics of phase transition in the new inflationary universe
scenario and generation of perturbations. Phys Lett B 117:175–178. doi:10.1016/0370-
2693(82)90541-X
79. Sylos LF, Vasilyev NL, Pietronero L, Baryshev YV (2009) Absence of self-averaging and of
homogeneity in the large-scale galaxy distribution. In: [40], p 49001. doi:10.1209/0295-5075/
86/49001. arXiv:0805.1132
80. Tomita K (2000) Distances and lensing in cosmological void models. ApJ 529:38–46. doi:10.
1086/308277. arXiv:astro-ph/9906027
81. Verde L, Protopapas P, Jimenez R (2013) Planck and the local Universe: quantifying the tension.
Phys Dark Universe 2:166–175. doi:10.1016/j.dark.2013.09.002. arXiv:1306.6766
82. Walker AG (1937) On milne’s theory of world-structure. Proc Lond Math Soc S2 42(1):90–127.
doi:10.1112/plms/s2-42.1.90. http://plms.oxfordjournals.org/content/s2-42/1/90.short. http://
plms.oxfordjournals.org/content/s2-42/1/90.full.pdf+html
83. Wang FY, Dai ZG (2013) Testing the local-void alternative to dark energy using galaxy pairs.
MNRAS 432:3025–3029. doi:10.1093/mnras/stt652. arXiv:1304.4399
84. Wu K, Lahav O, Rees M (1999) The large-scale smoothness of the Universe. Nature 397:225.
doi:10.1038/16637. arXiv:astro-ph/9804062
85. Yadav APS, Wandelt BD (2010) Primordial non-gaussianity in the cosmic microwave back-
ground. Adv Astron. doi:10.1155/2010/565248. arXiv:1006.0275
86. Yoo CM, Nakao K, Sasaki M (2010) CMB observations in LTB universes. Part II: the kSZ
effect in an LTB universe. J Cosmol Astropart Phys 10:011. doi:10.1088/1475-7516/2010/10/
011. arXiv:1008.0469
52 2 The Standard Model of Cosmology
87. York DG et al (2000) The sloan digital sky survey: technical summary. Astrophys J 120:1579–
1587. doi:10.1086/301513. arXiv:astro-ph/0006396
88. Zhang P, Stebbins A (2011) Confirmation of the Copernican principle at Gpc radial scale
and above from the kinetic Sunyaev-Zel’dovich effect power spectrum. Phys Rev Lett
107(4):041301. doi:10.1103/PhysRevLett.107.041301. arXiv:1009.3967
89. Zumalacárregui M, García-Bellido J, Ruiz-Lapuente P (2012) Tension in the void: cosmic rulers
strain inhomogeneous cosmologies. J Cosmol Astropart Phys 10:009. doi:10.1088/1475-7516/
2012/10/009. arXiv:1201.2790
http://www.springer.com/978-3-319-21881-6