0% found this document useful (0 votes)
21 views

Chapter1 Baumann Geometry Dynamics

Chapter1_Baumann_Geometry_Dynamics

Uploaded by

deboraalves2
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
21 views

Chapter1 Baumann Geometry Dynamics

Chapter1_Baumann_Geometry_Dynamics

Uploaded by

deboraalves2
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 23

1 Geometry and Dynamics

The further out we look into the universe, the simpler it seems to get (see Fig. 1.1). Averaged
over large scales, the clumpy distribution of galaxies becomes homogeneous and isotropic, i.e. in-
dependent of position and direction. As we will see, in §1.1, homogeneity and isotropy single
out a unique form of the spacetime geometry of the universe. We will discuss how particles and
light propagate in this spacetime in §1.2. Finally, in §1.3, we will show how the equations of
general relativity relate the rate of expansion of the universe to its matter content.

Figure 1.1: The distribution of galaxies is clumpy on small scales, but becomes more uniform on large scales
and at early times.

1.1 Geometry
1.1.1 Metric
I am assuming you have seen a metric before. (Otherwise, we will be in trouble.) Just to remind
you, the metric is an object that turns coordinate distances into physical distances. For example,
in three-dimensional Euclidean space, the physical distance between two points separated by the
infinitesimal coordinate distances dx, dy and dz is
3
X
2 2 2 2 i
d` = dx + dy + dz = ij dx dxj , (1.1.1)
ij=1

3
4 1. Geometry and Dynamics

where we have introduced the notation (x1 , x2 , x3 ) = (x, y, z). In this simple example, the metric
is the Kronecker delta ij = diag(1, 1, 1). However, you also know that if we were to use spherical
polar coordinates instead, the square of the physical distance would no longer be the sum of the
square of the coordinate distances. Instead, we would get
3
X
2 2 2 2 2 2 2
d` = dr + r d✓ + r sin ✓ d ⌘ gij dxi dxj , (1.1.2)
ij=1

where (x1 , x2 , x3 ) = (r, ✓, ). In this case, the metric has taken a less trivial form, namely
gij = diag(1, r2 , r2 sin2 ✓). This illustrates that observers using di↵erent coordinate systems
won’t necessarily agree on the coordinate distances between two points, but they will always
agree on the physical distance, d`. We say that d` is an invariant. Hence, the metric turns
observer-dependent coordinates into invariants.
A fundamental object in relativity is the spacetime metric. It turns observer-dependent
spacetime coordinates X µ = (t, xi ) into the invariant line element1
3
X
ds2 = gµ⌫ dX µ dX ⌫ ⌘ gµ⌫ dX µ dX ⌫ . (1.1.3)
µ,⌫=0

In special relativity, the Minkowski metric is the same everywhere in space and time,

gµ⌫ = diag(1, 1, 1, 1) . (1.1.4)

In general relativity, on the other hand, the metric will depend on the spacetime location,

gµ⌫ (t, x) . (1.1.5)

The spacetime dependence of the metric incorporates the e↵ects of gravity. How the metric
depends on the position in spacetime is determined by the distribution of matter and energy in
the universe. The large degree of symmetry of the homogeneous universe means that the metric
of the expanding universe take a rather simple form.
Spatial homogeneity and isotropy mean that the universe can be represented by a time-ordered
sequence of three-dimensional spatial slices ⌃t , each of which is homogeneous and isotropic (see
Fig. 1.2). The four-dimensional line element can then be written as

ds2 = dt2 a2 (t)d`2 , (1.1.6)

where d`2 ⌘ ij dxi dxj is the line element of a maximally symmetric 3-space and the scale factor
a(t) describes the expansion of the universe.

1.1.2 Symmetric Three-Spaces


We start with a classification of maximally symmetric 3-spaces. First, we note that homoge-
neous and isotropic 3-spaces have constant 3-curvature.2 There are only three options: zero
1
Throughout the course, will use the Einstein summation convention where repeated indices are summed
over. We will also use natural units with c ⌘ 1, so that dX 0 = dt. Our metric signature will be mostly
minus, (+, , , ).
2
We give a precise definition of Riemann curvature below.
5 1. Geometry and Dynamics

flat

negatively
curved

positively
curved

Figure 1.2: The spacetime of the universe can be foliated into flat, positively curved or negatively curved
spatial hypersurfaces.

curvature (E3 ), positive curvature (S3 ) and negative curvature (H3 ). The corresponding line
elements are
8
>
> 0 E3
dr2 <
2 2 2 2 2
d` = + r (d✓ + sin ✓ d ) , where k = +1 S3 . (1.1.7)
1 k r2 | {z } >
>
:
⌘ d⌦2 1 H3

Derivation⇤ .—Let us determine the metric for each case:


• flat space: the line element of three-dimensional Euclidean space E3 is simply

d`2 = dx2 = ij dx
i
dxj . (1.1.8)

This is clearly invariant under spatial translations (xi 7! xi +ai , with ai = const.) and rotations
(xi 7! Ri k xk , with ij Ri k Rj l = kl ).
• positively curved space: a 3-space with constant positive curvature can be represented as a
3-sphere S 3 embedded in four-dimensional Euclidean space E 4 ,

d`2 = dx2 + du2 , x2 + u2 = a2 , (1.1.9)

where a is the radius of the 3-sphere. Homogeneity and isotropy of the surface of the 3-sphere
are inherited from the symmetry of the line element under four-dimensional rotations.
• negatively curved space: a 3-space with constant negative curvature can be represented as a
hyperboloid H 3 embedded in four-dimensional Lorentzian space R1,3 ,

d`2 = dx2 du2 , x2 u2 = a2 , (1.1.10)

where a2 is an arbitrary constant. Homogeneity and isotropy of the induced geometry on


the hyperboloid are inherited from the symmetry of the line element under four-dimensional
pseudo-rotations (i.e. Lorentz transformations, with u playing the role of time).
In the last two cases, it is convenient to rescale the coordinates, x ! ax and u ! au. The line
elements of the spherical and hyperbolic cases then are
⇥ ⇤
d`2 = a2 dx2 ± du2 , x2 ± u2 = ±1 . (1.1.11)
6 1. Geometry and Dynamics

Notice that the coordinates x and u are now dimensionless, while the parameter a carries the di-
mension of length. The di↵erential of the embedding condition, x2 ± u2 = ±1, gives u du = ⌥ x · dx,
so 
(x · dx)2
d`2 = a2 dx2 ± . (1.1.12)
1 ⌥ x2
We can unify (1.1.12) with the Euclidean line element (1.1.8) by writing
8
 2 < 0 E3
(x · dx)
d`2 = a2 dx2 + k , for k⌘ +1 S3 . (1.1.13)
1 kx2 :
1 H3

Note that we must take a2 > 0 in order to have d`2 positive at x = 0, and hence everywhere. It is
convenient to use spherical polar coordinates, (r, ✓, ), because it makes the symmetries of the space
manifest. Using

dx2 = dr2 + r2 (d✓2 + sin2 ✓ d 2


), (1.1.14)
x · dx = r dr , (1.1.15)

the metric in (1.1.13) becomes diagonal



2 2 dr2
d` = a + r2 d⌦2 , (1.1.16)
1 k r2

where d⌦2 ⌘ d✓2 + sin2 ✓ d 2


.

Exercise.—Show that despite appearance r = 0 is not a special point in (1.1.7).

1.1.3 Robertson-Walker Metric


Substituting (1.1.7) into (1.1.6), we obtain the Robertson-Walker metric 3 in polar coordinates:

2 2 2 dr2
ds = dt a (t) + r2 d⌦2 . (1.1.17)
1 kr2

Notice that the symmetries of the universe have reduced the ten independent components of the
spacetime metric to a single function of time, the scale factor a(t), and a constant, the curvature
parameter k.

• The line element (1.1.17) has a rescaling symmetry


2
a ! a, r ! r/ , k! k. (1.1.18)

This means that the geometry of the spacetime stays the same if we simultaneously rescale
a, r and k as in (1.1.18). We can use this freedom to set the scale factor to unity today:4
a(t0 ) ⌘ 1. In this case, a(t) becomes dimensionless, and r and k 1/2 inherit the dimension
of length.
3
Sometimes this is called the Friedmann-Robertson-Walker (FRW) metric.
4
Quantities that are evaluated at the present time t0 will have a subscript ‘0’.
7 1. Geometry and Dynamics

time
Figure 1.3: Expansion of the universe. The comoving distance between points on an imaginary coordinate
grid remains constant as the universe expands. The physical distance is proportional to the comoving
distance times the scale factor a(t) and hence gets larger as time evolves.

• The coordinate r is called a comoving coordinate. Physical results depend only on the
physical coordinate rphys = a(t)r (see Fig. 1.3). The physical velocity of an object is
drphys dr da
vphys ⌘ = a(t) + r ⌘ vpec + Hrphys . (1.1.19)
dt dt dt
We see that this has two contributions: the so-called peculiar velocity, vpec ⌘ a(t) ṙ, and
the Hubble flow, Hrphys , where we have defined the Hubble parameter as 5

H⌘ . (1.1.20)
a
The peculiar velocity of an object is the velocity measured by a comoving observer (i.e. an
observer who follows the Hubble flow).

• The complicated grr component of (1.1.17) can sometimes be inconvenient. In that case,
p
we may redefine the radial coordinate, d ⌘ dr/ 1 kr2 , such that
⇥ ⇤
ds2 = dt2 a2 (t) d 2 + Sk2 ( ) d⌦2 , (1.1.21)

where 8 p
>
> sinh( k ) k<0
1 < p
Sk ( ) ⌘ p k k=0 . (1.1.22)
k>
>
: p
sin( k ) k>0
• It is also often useful to introduce conformal time,
dt
d⌘ = , (1.1.23)
a(t)
so that (1.1.21) becomes
h i
ds2 = a2 (⌘) d⌘ 2 d 2
+ Sk2 ( )d⌦2 . (1.1.24)

We see that the metric has factorized into a static metric multiplied by a time-dependent
conformal factor a(⌘). This form of the metric is particularly convenient for studying the
propagation of light.
5
Here, and in the following, an overdot denotes a time derivative, i.e. ȧ ⌘ da/dt.
8 1. Geometry and Dynamics

1.2 Kinematics
1.2.1 Geodesics
In this section, we will study how particles evolve in the FRW spacetime. Let us first, how-
ever, look at the simpler problem of the Newtonian dynamics of a free particle. In Cartesian
coordinates, we would simply have
d2 xi
= 0. (1.2.25)
dt2
We want to know what this equation turns into for an arbitrary coordinate system in which the
three-dimensional metric is gij 6= ij . To derive the equation of motion in these coordinates, we
start from the Lagrangian of the free particle
m
L= gij (xk ) ẋi ẋj . (1.2.26)
2
Substituting this into the Euler-Lagrange equation (see below), we find

d2 xi i dx
a
dxb
= ab , (1.2.27)
dt2 dt dt
where we have introduced the Christo↵el symbol

i 1
ab ⌘ g ij (@a gjb + @b gaj @j gab ) , with @j ⌘ @/@xj . (1.2.28)
2

Derivation.—The Euler-Lagrange equation is


✓ ◆
d @L @L
= . (1.2.29)
dt @ ẋk @xk

Substituting (1.2.26) into the derivatives, we have

@L 1
k
= @k gij ẋi ẋj , (1.2.30)
@x 2
@L
= gik ẋi , (1.2.31)
@ ẋk
where we have set m ⌘ 1 since it will cancel on both sides. The l.h.s. of (A.2.32) then becomes
✓ ◆
d @L d i i dxj @gik i
= g ik ẋ = g ik ẍ + ẋ
dt @ ẋk dt dt @xj
= gik ẍi + @j gik ẋi ẋj
1
= gik ẍi + (@i gjk + @j gik ) ẋi ẋj . (1.2.32)
2
Equation (A.2.32) then implies
1
gki ẍi = (@i gjk + @j gik @k gij ) ẋi ẋj . (1.2.33)
2
Multiplying both sides by g lk , we get
1 lk
ẍl = g (@i gjk + @j gik @k gij ) ẋi ẋj ⌘ l i j
ij ẋ ẋ , (1.2.34)
2
which is the desired result (1.2.27).
9 1. Geometry and Dynamics

The equation of motion of a massive particle in general relativity will take a similar form as
eq. (1.2.27). However, in this case, the term involving the Christo↵el symbol cannot be re-
moved by going to Cartesian coordinates, but is a physical manifestation of the curvature of the
spacetime.

Geodesic Equation⇤

In the absence of additional non-gravitational forces, freely-falling particles in a curved spacetime


move along special trajectories called geodesics. For massive particles, a geodesic is the timelike
curve X µ (⌧ ) which extremises the proper time ⌧ between two points in the spacetime. In
Appendix A, I show that this extremal path satisfies the geodesic equation

d2 X µ µ dX ↵ dX
= ↵ , (1.2.35)
d⌧ 2 d⌧ d⌧

where
µ 1
↵⌘ g µ (@↵ g + @ g↵ @ g↵ ) . (1.2.36)
2
Notice the similarity between eqs. (1.2.35) and (1.2.27). It will be convenient to write the
geodesic equation in a few di↵erent ways:

• Introducing the four-velocity of the particle, U µ ⌘ dX µ /d⌧ , we get

dU µ µ
= ↵ U ↵U . (1.2.37)
d⌧
Using the chain rule
d µ ↵ dX ↵ @U µ ↵ @U
µ
U (X (⌧ )) = = U , (1.2.38)
d⌧ d⌧ @X ↵ @X ↵
we can also write this as ✓ ◆
@U µ µ
U↵ + ↵ U = 0. (1.2.39)
@X ↵
The term in brackets is the covariant derivative of the four-vector U µ , i.e. r↵ U µ ⌘ @↵ U µ +
µ
↵ U . This allows us to write the geodesic equation in the following slick way:

U ↵ r↵ U µ = 0 . (1.2.40)

In the GR course you will derive this form of the geodesic equation directly by thinking
about parallel transport.

• Using the definition of the four-momentum of the massive particle, P µ = mU µ , we can


write (1.2.40) as

@P µ µ
P ↵ r↵ P µ = 0 or P↵ = ↵ P ↵P . (1.2.41)
@X ↵

This form of the geodesic equation is useful since it also applies to massless particles.

I will now show you how to apply the geodesic equation (1.2.41) to particles in the FRW universe.
10 1. Geometry and Dynamics

Particles in the Expanding Universe

To evaluate the r.h.s. of (1.2.41) we need to compute the Christo↵el symbols for the FRW metric,

ds2 = dt2 a2 (t) ij dx


i
dxj . (1.2.42)
µ 0
All Christo↵el symbols with two time indices vanish, i.e. 00 = 0 = 0. The only non-zero
components are

0 i ȧ i i 1 il
ij = aȧ ij , 0j = j , jk = (@j kl + @k jl @l jk ) , (1.2.43)
a 2
µ µ 0
or are related to these by symmetry (note that ↵ = ↵ ). I will derive ij as an example and
leave i0j as an exercise.

Example.—The Christo↵el symbol with upper index equal to zero is

0 1 0
↵ = g (@↵ g + @ g↵ @ g↵ ) . (1.2.44)
2
The factor g 0 vanishes unless = 0 in which case it is equal to 1. Therefore,

0 1
↵ = (@↵ g 0 + @ g↵0 @0 g↵ ) . (1.2.45)
2
The first two terms reduce to derivatives of g00 (since gi0 = 0). The FRW metric has constant g00 ,
so these terms vanish and we are left with

0 1
↵ = @0 g↵ . (1.2.46)
2
The derivative is non-zero only if ↵ and are spatial indices, gij = a2 ij (don’t miss the sign!). In
that case, we find
0
ij = aȧ ij . (1.2.47)

The homogeneity of the FRW background implies @i P µ = 0, so that the geodesic equation (1.2.41)
reduces to
dP µ µ
P0 = ↵ P ↵P ,
dt ⇣ ⌘
µ 0 µ i
= 2 0j P + ij P Pj , (1.2.48)

where I have used (1.2.43) in the second line.

• The first thing to notice from (1.2.48) is that massive particles at rest in the comoving
frame, P j = 0, will stay at rest because the r.h.s. then vanishes,

dP i
Pj = 0 ) = 0. (1.2.49)
dt

• Next, we consider the µ = 0 component of (1.2.48), but don’t require the particles to be
at rest. The first term on the r.h.s. vanishes because 00j = 0. Using (1.2.43), we then find

dE 0 i j ȧ 2
E = ij P P = p , (1.2.50)
dt a
11 1. Geometry and Dynamics

where we have written P 0 ⌘ E and defined the amplitude of the physical three-momentum
as
p2 ⌘ gij P i P j = a2 ij P i P j . (1.2.51)
Notice the appearance of the scale factor in (1.2.51) from the contraction with the spatial
part of the FRW metric, gij = a2 ij . The components of the four-momentum satisfy
the constraint gµ⌫ P µ P ⌫ = m2 , or E 2 p2 = m2 , where the r.h.s. vanishes for massless
particles. It follows that E dE = pdp, so that (1.2.50) can be written as

ṗ ȧ 1
= ) p/ . (1.2.52)
p a a

We see that the physical three-momentum of any particle (both massive and massless)
decays with the expansion of the universe.

– For massless particles, eq. (1.2.52) implies


1
p=E/ (massless particles) , (1.2.53)
a
i.e. the energy of massless particles decays with the expansion.
– For massive particles, eq. (1.2.52) implies
mv 1
p= p / (massive particles) , (1.2.54)
1 v 2 a

where v i = dxi /dt is the comoving peculiar velocity of the particles (i.e. the velocity
relative to the comoving frame) and v 2 ⌘ a2 ij v i v j is the magnitude of the physical
peculiar velocity, cf. eq. (1.1.19). To get the first equality in (1.2.54), I have used

dX i dt mv i mv i
P i = mU i = m = m vi = p =p . (1.2.55)
d⌧ d⌧ 1 a2 ij v i v j 1 v2

Equation (1.2.54) shows that freely-falling particles left on their own will converge
onto the Hubble flow.

1.2.2 Redshift
Everything we know about the universe is inferred from the light we receive from distant ob-
jects. The light emitted by a distant galaxy can be viewed either quantum mechanically as
freely-propagating photons, or classically as propagating electromagnetic waves. To interpret
the observations correctly, we need to take into account that the wavelength of the light gets
stretched (or, equivalently, the photons lose energy) by the expansion of the universe. We now
quantify this e↵ect.
Photons.—In the quantum mechanical description, the wavelength of light is inversely propor-
tional to the photon momentum, = h/p. Since according to (1.2.53) the momentum of a photon
evolves as a(t) 1 , the wavelength scales as a(t). Light emitted at time t1 with wavelength 1
will be observed at t0 with wavelength

a(t0 )
0 = 1 . (1.2.56)
a(t1 )
12 1. Geometry and Dynamics

Since a(t0 ) > a(t1 ), the wavelength of the light increases, 0 > 1.

Classical waves.—We can derive the same result by treating light as classical electromagnetic
waves. Consider a galaxy at a fixed comoving distance d. At a time ⌘1 , the galaxy emits a signal
of short conformal duration ⌘. The light arrives at our telescopes at time ⌘0 = ⌘1 + d. The
conformal duration of the signal measured by the detector is the same as at the source, but the
physical time intervals are di↵erent at the points of emission and detection,

t1 = a(⌘1 ) ⌘ and t0 = a(⌘0 ) ⌘ . (1.2.57)

If t is the period of the light wave, the light is emitted with wavelength 1 = t1 (in units
where c = 1), but is observed with wavelength 0 = t0 , so that
0 a(⌘0 )
= . (1.2.58)
1 a(⌘1 )

Redshift.—It is conventional to define the redshift parameter as the fractional shift in wavelength
of a photon emitted by a distant galaxy at time t1 and observed on Earth today,
0 1
z⌘ . (1.2.59)
1

We then find
a(t0 )
1+z = . (1.2.60)
a(t1 )
It is also common to define a(t0 ) ⌘ 1, so that

1
1+z = . (1.2.61)
a(t1 )

Hubble’s law.—For nearby sources, we may expand a(t1 ) in a power series,


⇥ ⇤
a(t1 ) = a(t0 ) 1 + (t1 t0 )H0 + · · · , (1.2.62)

where H0 is the Hubble constant


ȧ(t0 )
H0 ⌘ . (1.2.63)
a(t0 )
Equation (1.2.60) then gives z = H0 (t0 t1 ) + · · · . For close objects, t0 t1 is simply the
physical distance d (in units with c = 1). We therefore find that the redshift increases linearly
with distance
z ' H0 d . (1.2.64)
The slope in a redshift-distance diagram (cf. fig. 1.6) therefore measures the current expansion
rate of the universe, H0 . These measurements used to come with very large uncertainties. Since
H0 normalizes everything else (see below), it became conventional to define6
1 1
H0 ⌘ 100h kms Mpc , (1.2.65)

where the parameter h is used to keep track of how uncertainties in H0 propagate into other
cosmological parameters. Today, measurements of H0 have become much more precise,7

h ⇡ 0.67 ± 0.01 . (1.2.66)


6
A parsec (pc) is 3.26 light-years. Blame astronomers for the funny units in (6.3.29).
7
Planck 2015 Results: Cosmological Parameters [arXiv:1502.01589].
13 1. Geometry and Dynamics

1.2.3 Distances⇤
For distant objects, we have to be more careful about what we mean by “distance”:

• Metric distance.—We first define a distance that isn’t really observable, but that will be
useful in defining observable distances. Consider the FRW metric in the form (1.1.21),
h i
ds2 = dt2 a2 (t) d 2 + Sk2 ( )d⌦2 , (1.2.67)

where8 8
>
< R0 sinh( /R0 ) k= 1
Sk ( ) ⌘ k=0 . (1.2.68)
>
: R sin( /R )
0 0 k = +1
The distance multiplying the solid angle element d⌦2 is the metric distance,

d m = Sk ( ) . (1.2.69)

In a flat universe (k = 0), the metric distance is simply equal to the comoving distance .
The comoving distance between us and a galaxy at redshift z can be written as
Z t0 Z z
dt dz
(z) = = , (1.2.70)
t1 a(t) 0 H(z)

where the redshift evolution of the Hubble parameter, H(z), depends on the matter content
of the universe (see §1.3). We emphasize that the comoving distance and the metric
distance are not observables.

• Luminosity distance.—Type IA supernovae are called ‘standard candles’ because they are
believed to be objects of known absolute luminosity L (= energy emitted per second).
The observed flux F (= energy per second per receiving area) from a supernova explosion
can then be used to infer its (luminosity) distance. Consider a source at a fixed comoving
distance . In a static Euclidean space, the relation between absolute luminosity and
observed flux is
L
F = . (1.2.71)
4⇡ 2

In an FRW spacetime, this result is modified for three reasons:

1. At the time t0 that the light reaches the Earth, the proper area of a sphere drawn
around the supernova and passing through the Earth is 4⇡d2m . The fraction of the
light received in a telescope of aperture A is therefore A/4⇡d2m .
2. The rate of arrival of photons is lower than the rate at which they are emitted by the
redshift factor 1/(1 + z).
3. The energy E0 of the photons when they are received is less than the energy E1 with
which they were emitted by the same redshift factor 1/(1 + z).
8
Notice that the definition of Sk ( ) contains a length scale R0 after we chose to make the scale factor dimen-
sionless, a(t0 ) ⌘ 1. This is achieved by using the rescaling symmetry a ! a, ! / , and Sk2 ! Sk2 / .
14 1. Geometry and Dynamics

observer

source

Figure 1.4: Geometry associated with the definition of luminosity distance.

Hence, the correct formula for the observed flux of a source with luminosity L at coordinate
distance and redshift z is
L L
F = ⌘ , (1.2.72)
4⇡d2m (1 + z)2 4⇡d2L

where we have defined the luminosity distance, dL , so that the relation between luminosity,
flux and luminosity distance is the same as in (1.2.71). Hence, we find

dL = dm (1 + z) . (1.2.73)

• Angular diameter distance.—Sometimes we can make use of ‘standard rulers’, i.e. objects
of known physical size D. (This is the case, for example, for the fluctuations in the CMB.)
Let us assume again that the object is at a comoving distance and the photons which
we observe today were emitted at time t1 . A naive astronomer could decide to measure
the distance dA to the object by measuring its angular size ✓ and using the Euclidean
formula for its distance,9
D
dA = . (1.2.74)

This quantity is called the angular diameter distance. The FRW metric (1.1.24) implies

observer

source

Figure 1.5: Geometry associated with the definition of angular diameter distance.

the following relation between the physical (transverse) size of the object and its angular
size on the sky
dm
D = a(t1 )Sk ( ) ✓ = ✓. (1.2.75)
1+z
9
This formula assumes ✓ ⌧ 1 (in radians) which is true for all cosmological objects.
15 1. Geometry and Dynamics

Hence, we get
dm
.dA = (1.2.76)
1+z
The angular diameter distance measures the distance between us and the object when
the light was emitted. We see that angular diameter and luminosity distances aren’t
independent, but related by
dL
dA = . (1.2.77)
(1 + z)2
Fig. 1.6 shows the redshift dependence of the three distance measures dm , dL , and dA . Notice
that all three distances are larger in a universe with dark energy (in the form of a cosmological
constant ⇤) than in one without. This fact was employed in the discovery of dark energy (see
fig. 1.7 in §1.3.3).

with

without
distance

redshift
Figure 1.6: Distance measures in a flat universe, with matter only (dotted lines) and with 70% dark energy
(solid lines). In a dark energy dominated universe, distances out to a fixed redshift are larger than in a
matter-dominated universe.

1.3 Dynamics
The dynamics of the universe is determined by the Einstein equation
Gµ⌫ = 8⇡GTµ⌫ . (1.3.78)
This relates the Einstein tensor Gµ⌫ (a measure of the “spacetime curvature” of the FRW
universe) to the stress-energy tensor Tµ⌫ (a measure of the “matter content” of the universe). We
will first discuss possible forms of cosmological stress-energy tensors Tµ⌫ (§1.3.1), then compute
the Einstein tensor Gµ⌫ for the FRW background (§1.3.2), and finally put them together to solve
for the evolution of the scale factor a(t) as a function of the matter content (§1.3.3).

1.3.1 Matter Sources


We first show that the requirements of isotropy and homogeneity force the coarse-grained stress-
energy tensor to be that of a perfect fluid,
Tµ⌫ = (⇢ + P ) Uµ U⌫ P gµ⌫ , (1.3.79)
where ⇢ and P are the energy density and the pressure of the fluid and U µ is its four-velocity
(relative to the observer).
16 1. Geometry and Dynamics

Number Density

In fact, before we get to the stress-energy tensor, we study a simpler object: the number current
four-vector N µ . The µ = 0 component, N 0 , measures the number density of particles, where for
us a “particle” may be an entire galaxy. The µ = i component, N i , is the flux of the particles in
the direction xi . Isotropy requires that the mean value of any 3-vector, such as N i , must vanish,
and homogeneity requires that the mean value of any 3-scalar10 , such as N 0 , is a function only
of time. Hence, the current of galaxies, as measured by a comoving observer, has the following
components
N 0 = n(t) , Ni = 0 , (1.3.80)
where n(t) is the number of galaxies per proper volume as measured by a comoving observer.
A general observer (i.e. an observer in motion relative to the mean rest frame of the particles),
would measure the following number current four-vector

N µ = nU µ , (1.3.81)

where U µ ⌘ dX µ /d⌧ is the relative four-velocity between the particles and the observer. Of
course, we recover the previous result (1.3.80) for a comoving observer, U µ = (1, 0, 0, 0). For
U µ = (1, v i ), eq. (1.3.81) gives the correctly boosted results. For instance, you may recall that
the boosted number density is n. (The number density increases because one of the dimensions
of the volume is Lorentz contracted.)
The number of particles has to be conserved. In Minkowski space, this implies that the
evolution of the number density satisfies the continuity equation

Ṅ 0 = @i N i , (1.3.82)

or, in relativistic notation,


@µ N µ = 0 . (1.3.83)
Equation (1.3.83) is generalised to curved spacetimes by replacing the partial derivative @µ with
a covariant derivative rµ ,11
rµ N µ = @µ N µ + µµ N = 0 . (1.3.84)
Using (1.3.80), this becomes
dn i
+ i0 n = 0, (1.3.85)
dt
and substituting (1.2.43), we find

ṅ ȧ 3
= 3 ) n(t) / a . (1.3.86)
n a
As expected, the number density decreases in proportion to the increase of the proper volume.
10
A 3-scalar is a quantity that is invariant under purely spatial coordinate transformations.
11
The covariant derivative is an important object in di↵erential geometry and it is of fundamental importance
in general relativity. The geometrical meaning of rµ will be discussed in detail in the GR course. In this course,
we will have to be satisfied with treating it as an operator that acts in a specific way on scalars, vectors and
tensors.
17 1. Geometry and Dynamics

Energy-Momentum Tensor

We will now use a similar logic to determine what form of the stress-energy tensor Tµ⌫ is
consistent with the requirements of homogeneity and isotropy. First, we decompose Tµ⌫ into a
3-scalar, T00 , 3-vectors, Ti0 and T0j , and a 3-tensor, Tij . As before, isotropy requires the mean
values of 3-vectors to vanish, i.e. Ti0 = T0j = 0. Moreover, isotropy around a point x = 0
requires the mean value of any 3-tensor, such as Tij , at that point to be proportional to ij and
hence to gij , which equals a2 ij at x = 0,

Tij (x = 0) / ij / gij (x = 0) . (1.3.87)

Homogeneity requires the proportionality coefficient to be only a function of time. Since this is
a proportionality between two 3-tensors, Tij and gij , it must remain una↵ected by an arbitrary
transformation of the spatial coordinates, including those transformations that preserve the form
of gij while taking the origin into any other point. Hence, homogeneity and isotropy require the
components of the stress-energy tensor everywhere to take the form

T00 = ⇢(t) , ⇡i ⌘ Ti0 = 0 , Tij = P (t)gij (t, x) . (1.3.88)

It looks even nicer with mixed upper and lower indices


0 1
⇢ 0 0 0
B 0 P 0 0 C
B C
T µ⌫ = gµ T ⌫ = B C. (1.3.89)
@ 0 0 P 0 A
0 0 0 P

This is the stress-energy tensor of a perfect fluid as seen by a comoving observer. More generally,
the stress-energy tensor can be written in the following, explicitly covariant, form

T µ ⌫ = (⇢ + P ) U µ U⌫ P µ
⌫ , (1.3.90)

where U µ ⌘ dX µ /d⌧ is the relative four-velocity between the fluid and the observer, while ⇢ and
P are the energy density and pressure in the rest-frame of the fluid. Of course, we recover the
previous result (1.3.89) for a comoving observer, U µ = (1, 0, 0, 0).
How do the density and pressure evolve with time? In Minkowski space, energy and momen-
tum are conserved. The energy density therefore satisfies the continuity equation ⇢˙ = @i ⇡ i ,
i.e. the rate of change of the density equals the divergence of the energy flux. Similarly, the
evolution of the momentum density satisfies the Euler equation, ⇡˙ i = @i P . These conservation
laws can be combined into a four-component conservation equation for the stress-energy tensor

@µ T µ ⌫ = 0 . (1.3.91)

In general relativity, this is promoted to the covariant conservation equation


µ
rµ T µ ⌫ = @ µ T µ ⌫ + µ T ⌫ µ⌫ T
µ
= 0. (1.3.92)

This corresponds to four separate equations (one for each ⌫). The evolution of the energy density
is determined by the ⌫ = 0 equation
µ
@µ T µ 0 + µ T 0 µ0 T
µ
= 0. (1.3.93)
18 1. Geometry and Dynamics

Since T i 0 vanishes by isotropy, this reduces to


d⇢ µ µ
+ µ0 ⇢ µ0 T = 0. (1.3.94)
dt
From eq. (1.2.43) we see that µ0 vanishes unless and µ are spatial indices equal to each other,
in which case it is ȧ/a. The continuity equation (1.3.94) therefore reads


⇢˙ + 3 (⇢ + P ) = 0 . (1.3.95)
a

Exercise.—Show that (1.3.95) can be written as, dU = P dV , where U = ⇢V and V / a3 .

Cosmic Inventory

The universe is filled with a mixture of di↵erent matter components. It is useful to classify the
di↵erent sources by their contribution to the pressure:

• Matter
We will use the term “matter” to refer to all forms of matter for which the pressure is
much smaller than the energy density, |P | ⌧ ⇢. As we will show in Chapter 3, this is the
case for a gas of non-relativistic particles (where the energy density is dominated by the
mass). Setting P = 0 in (1.3.95) gives
3
⇢/a . (1.3.96)

This dilution of the energy density simply reflects the expansion of the volume V / a3 .

– Dark matter. Most of the matter in the universe is in the form of invisible dark
matter. This is usually thought to be a new heavy particle species, but what it really
is, we don’t know.
– Baryons. Cosmologists refer to ordinary matter (nuclei and electrons) as baryons.12

• Radiation
We will use the term “radiation” to denote anything for which the pressure is about a
third of the energy density, P = 13 ⇢. This is the case for a gas of relativistic particles, for
which the energy density is dominated by the kinetic energy (i.e. the momentum is much
bigger than the mass). In this case, eq. (1.3.95) implies
4
⇢/a . (1.3.97)

The dilution now includes the redshifting of the energy, E / a 1.

– Photons. The early universe was dominated by photons. Being massless, they are al-
ways relativistic. Today, we detect those photons in the form of the cosmic microwave
background.
12
Of course, this is technically incorrect (electrons are leptons), but nuclei are so much heavier than electrons
that most of the mass is in the baryons. If this terminology upsets you, you should ask your astronomer friends
what they mean by “metals”.
19 1. Geometry and Dynamics

– Neutrinos. For most of the history of the universe, neutrinos behaved like radiation.
Only recently have their small masses become relevant and they started to behave
like matter.
– Gravitons. The early universe may have produced a background of gravitons (i.e. grav-
itational waves, see §6.5.2). Experimental e↵orts are underway to detect them.

• Dark energy
We have recently learned that matter and radiation aren’t enough to describe the evolution
of the universe. Instead, the universe today seems to be dominated by a mysterious negative
pressure component, P = ⇢. This is unlike anything we have ever encountered in the
lab. In particular, from eq. (1.3.95), we find that the energy density is constant,

⇢ / a0 . (1.3.98)

Since the energy density doesn’t dilute, energy has to be created as the universe expands.13

– Vacuum energy. In quantum field theory, this e↵ect is actually predicted! The ground
state energy of the vacuum corresponds to the following stress-energy tensor
vac
Tµ⌫ = ⇢vac gµ⌫ . (1.3.99)

Comparison with eq. (1.3.90), show that this indeed implies Pvac = ⇢vac . Unfortu-
nately, the predicted size of ⇢vac is completely o↵,
⇢vac
⇠ 10120 . (1.3.100)
⇢obs
This so-called “cosmological constant problem” is the biggest crisis in modern theo-
retical physics.
– Something else? The failure of quantum field theory to explain the size of the observed
dark energy has lead theorists to consider more exotic possibilities (such as time-
varying dark energy and modifications of general relativity). In my opinion, none of
these ideas works very well.

Cosmological constant.—The left-hand side of the Einstein equation (1.3.78) isn’t uniquely defined.
We can add the term ⇤gµ⌫ , for some constant ⇤, without changing the conservation of the stress
tensor, rµ Tµ⌫ = 0 (recall, or check, that rµ gµ⌫ = 0). In other words, we could have written the
Einstein equation as
Gµ⌫ ⇤gµ⌫ = 8⇡GTµ⌫ . (1.3.101)
Einstein, in fact, did add such a term and called it the cosmological constant. However, it has become
modern practice to move this term to the r.h.s. and treat it as a contribution to the stress-energy
tensor of the form
(⇤) ⇤
Tµ⌫ = gµ⌫ ⌘ ⇢⇤ gµ⌫ . (1.3.102)
8⇡G
This is of the same form as the stress-energy tensor from vacuum energy, eq. (1.3.99).

13
In a gravitational system this doesn’t have to violate the conservation of energy. It is the conservation equation
(1.3.95) that counts.
20 1. Geometry and Dynamics

Summary

Most cosmological fluids can be parameterised in terms of a constant equation of state: w = P/⇢.
This includes cold dark matter (w = 0), radiation (w = 1/3) and vacuum energy (w = 1). In
that case, the solutions to (1.3.95) scale as
8
> 3
< a matter
3(1+w)
⇢/a = a 4 radiation . (1.3.103)
>
: a0 vacuum

1.3.2 Spacetime Curvature


We want to relate these matter sources to the evolution of the scale factor in the FRW metric.
To do this we have to compute the Einstein tensor on the l.h.s. of the Einstein equation (1.3.78),
1
Gµ⌫ = Rµ⌫ Rgµ⌫ . (1.3.104)
2
We will need the Ricci tensor
⇢ ⇢
Rµ⌫ ⌘ @ µ⌫ @⌫ µ + ⇢ µ⌫ µ ⌫⇢ , (1.3.105)

and the Ricci scalar


R = Rµ µ = g µ⌫ Rµ⌫ . (1.3.106)
Again, there is a lot of beautiful geometry behind these definitions. We will simply keep plugging-
and-playing: given the Christo↵el symbols (1.2.43) nothing stops us from computing (1.3.105).
We don’t need to calculate Ri0 = R0i , because it is a 3-vector, and therefore must vanish
due to the isotropy of the Robertson-Walker metric. (Try it, if you don’t believe it!) The
non-vanishing components of the Ricci tensor are

R00 = 3 , (1.3.107)
"a ✓ ◆2 #
ä ȧ k
Rij = +2 + 2 2 gij . (1.3.108)
a a a

Notice that we had to find Rij / gij to be consistent with homogeneity and isotropy.

Derivation of R00 .—Setting µ = ⌫ = 0 in (1.3.105), we have


⇢ ⇢
R00 = @ 00 @0 0 + ⇢ 00 0 0⇢ , (1.3.109)

Since Christo↵els with two time-components vanish, this reduces to


i i j
R00 = @0 0i 0j 0i . (1.3.110)
i
Using 0j = (ȧ/a) ji , we find
✓ ◆ ✓ ◆2
d ȧ ȧ ä
R00 = 3 3 = 3 . (1.3.111)
dt a a a
21 1. Geometry and Dynamics

The Ricci scalar is

R = g µ⌫ Rµ⌫
" ✓ ◆2 #
1 ä ȧ k
= R00 Rii = 6 + + 2 . (1.3.112)
a2 a a a

The non-zero components of the Einstein tensor Gµ ⌫ ⌘ g µ G ⌫ then are


"✓ ◆ #
0 ȧ 2 k
G 0 = 3 + 2 , (1.3.113)
a a
" ✓ ◆2 #
ä ȧ k
Gi j = 2 + + 2 ji . (1.3.114)
a a a

Exercise.—Verify eqs. (1.3.113) and (1.3.114).

1.3.3 Friedmann Equations


Combining eqs. (1.3.113) and (1.3.114) with the stress-tensor (1.3.89), we get the Friedmann
equations,
✓ ◆2
ȧ 8⇡G k
= ⇢ , (1.3.115)
a 3 a2
ä 4⇡G
= (⇢ + 3P ) . (1.3.116)
a 3
Here, ⇢ and P should be understood as the sum of all contributions to the energy density and
pressure in the universe. We write ⇢r for the contribution from radiation (with ⇢ for photons
and ⇢⌫ for neutrinos), ⇢m for the contribution by matter (with ⇢c for cold dark matter and ⇢b
for baryons) and ⇢⇤ for the vacuum energy contribution. The first Friedmann equation is often
written in terms of the Hubble parameter, H ⌘ ȧ/a,

8⇡G k
H2 = ⇢ . (1.3.117)
3 a2
Let us use subscripts ‘0’ to denote quantities evaluated today, at t = t0 . A flat universe (k = 0)
corresponds to the following critical density today
3H02 29
⇢crit,0 = = 1.9 ⇥ 10 h2 grams cm 3
8⇡G
= 2.8 ⇥ 1011 h2 M Mpc 3

5
= 1.1 ⇥ 10 h2 protons cm 3
. (1.3.118)

We use the critical density to define dimensionless density parameters


⇢I,0
⌦I,0 ⌘ . (1.3.119)
⇢crit,0
The Friedmann equation (1.3.117) can then be written as
 ⇣ a ⌘4 ⇣ a ⌘3 ⇣ a ⌘2
0 0 0
H 2 (a) = H02 ⌦r,0 + ⌦m,0 + ⌦k,0 + ⌦⇤,0 , (1.3.120)
a a a
22 1. Geometry and Dynamics

26 HST 0.32 0.68

distance (apparent magnitude)


1.00 0.00
24 SNLS

22 SDSS

20

18

16 Low-z

14

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4


redshift

Figure 1.7: Type IA supernovae and the discovery dark energy. If we assume a flat universe, then the
supernovae clearly appear fainter (or more distant) than predicted in a matter-only universe (⌦m = 1.0).
(SDSS = Sloan Digital Sky Survey; SNLS = SuperNova Legacy Survey; HST = Hubble Space Telescope.)

where we have defined a “curvature” density parameter, ⌦k,0 ⌘ k/(a0 H0 )2 . It should be noted
that in the literature, the subscript ‘0’ is normally dropped, so that e.g. ⌦m usually denotes
the matter density today in terms of the critical density today. From now on we will follow
this convention and drop the ‘0’ subscripts on the density parameters. We will also use the
conventional normalization for the scale factor, a0 ⌘ 1. Equation (1.3.120) then becomes
H2 4 3 2
= ⌦r a + ⌦m a + ⌦k a + ⌦⇤ . (1.3.121)
H02

⇤CDM

Observations (see figs. 1.7 and 1.8) show that the universe is filled with radiation (‘r’), matter
(‘m’) and dark energy (‘⇤’):
5
|⌦k |  0.01 , ⌦r = 9.4 ⇥ 10 , ⌦m = 0.32 , ⌦⇤ = 0.68 .

The equation of state of dark energy seems to be that of a cosmological constant, w⇤ ⇡ 1. The
matter splits into 5% ordinary matter (baryons, ‘b’) and 27% (cold) dark matter (CDM, ‘c’):

⌦b = 0.05 , ⌦c = 0.27 .

We see that even today curvature makes up less than 1% of the cosmic energy budget. At earlier
times, the e↵ects of curvature are then completely negligible (recall that matter and radiation
scale as a 3 and a 4 , respectively, while the curvature contribution only increases as a 2 ). For
the rest of these lectures, I will therefore set ⌦k ⌘ 0. In Chapter 2, we will show that inflation
indeed predicts that the e↵ects of curvature should be minuscule in the early universe (see also
Problem Set 2).

Single-Component Universe

The di↵erent scalings of radiation (a 4 ), matter (a 3 ) and vacuum energy (a0 ) imply that for
most of its history the universe was dominated by a single component (first radiation, then
23 1. Geometry and Dynamics

0.80
+lensing 75
+lensing+BAO
70
0.72
65

60
0.64
55

50
0.56
45

40
0.24 0.32 0.40 0.48

Figure 1.8: A combination CMB and LSS observations indicate that the spatial geometry of the universe
is flat. The energy density of the universe is dominated by a cosmological constant. Notice that the CMB
data alone cannot exclude a matter-only universe with large spatial curvature. The evidence for dark energy
requires additional input.

matter, then vacuum energy; see fig. 1.9). Parameterising this component by its equation of
state wI captures all cases of interest. For a flat, single-component universe, the Friedmann
equation (1.3.121) reduces to
ȧ p 3
= H0 ⌦I a 2 (1+wI ) . (1.3.122)
a
Integrating this equation, we obtain the time dependence of the scale factor
8
>
> t2/3 MD
>
> t 2/3(1+wI ) w 6
= 1
>
<
I
t1/2 RD
a(t) / (1.3.123)
>
>
>
>
>
: eHt wI = 1 ⇤D

matter

radiation

cosmological constant

Figure 1.9: Evolution of the energy densities in the universe.


24 1. Geometry and Dynamics

or, in conformal time,


8
>
> ⌘2 MD
>
> ⌘ 2/(1+3wI ) wI 6= 1
>
< ⌘ RD
a(⌘) / (1.3.124)
>
>
>
>
>
: ( ⌘) 1 wI = 1 ⇤D

Exercise.—Derive eq. (1.3.124) from eq. (1.3.123).

Two-Component Universe⇤

Matter and radiation were equally important at aeq ⌘ ⌦r /⌦m ⇡ 3 ⇥ 10 4 , which was shortly
before the cosmic microwave background was released (in §3.3.3, we will show that this happened
at arec ⇡ 9 ⇥ 10 4 ). It will be useful to have an exact solution describing the transition era. Let
us therefore consider a flat universe filled with a mixture of matter and radiation. To solve for
the evolution of the scale factor, it proves convenient to move to conformal time. The Friedmann
equations (1.3.115) and (1.3.116) then are
8⇡G 4
(a0 )2 = ⇢a , (1.3.125)
3
4⇡G
a00 = (⇢ 3P )a3 , (1.3.126)
3
where primes denote derivatives with respect to conformal time and

⇢eq ⇣ aeq ⌘3 ⇣ aeq ⌘4
⇢ ⌘ ⇢m + ⇢r = + . (1.3.127)
2 a a

Exercise.—Derive eqs. (1.3.125) and (1.3.126). You will first need to convince yourself that ȧ = a0 /a
and ä = a00 /a2 (a0 )2 /a3 .

Notice that radiation doesn’t contribute as a source term in eq. (1.3.126), ⇢r 3Pr = 0. Moreover,
since ⇢m a3 = const. = 12 ⇢eq a3eq , we can write eq. (1.3.126) as
2⇡G
a00 = ⇢eq a3eq . (1.3.128)
3
This equation has the following solution
⇡G
a(⌘) = ⇢eq a3eq ⌘ 2 + C⌘ + D . (1.3.129)
3
Imposing a(⌘ = 0) ⌘ 0, fixes one integration constant, D = 0. We find the second integration
constant by substituting (1.3.129) and (1.3.127) into (1.3.125),
✓ ◆1/2
4⇡G 4
C= ⇢eq aeq . (1.3.130)
3
Eq. (1.3.129) can then be written as
"✓ ◆2 ✓ ◆#
⌘ ⌘
a(⌘) = aeq +2 , (1.3.131)
⌘? ⌘?
25 1. Geometry and Dynamics

where ✓ ◆ 1/2
⇡G ⌘eq
⌘? ⌘ ⇢eq a2eq =p . (1.3.132)
3 2 1
For ⌘ ⌧ ⌘eq , we recover the radiation-dominated limit, a / ⌘, while for ⌘ ⌘eq , we agree with
the matter-dominated limit, a / ⌘ 2 .

You might also like