02.04.2014 Views

Chapter 3: THE FRIEDMANN MODELS

Chapter 3: THE FRIEDMANN MODELS

Chapter 3: THE FRIEDMANN MODELS

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>Chapter</strong> 3: <strong>THE</strong> <strong>FRIEDMANN</strong> <strong>MODELS</strong><br />

In this chapter we incorporate gravity and our knowledge of how the<br />

density of the Universe changes as it expands to generate the family of<br />

Friedmann models for R(τ). The dynamics and the curvature are both<br />

determined by the average density, parameterized by Ω. These<br />

models enable us to calculate the R 0 S k (ω) term in the Robertson-<br />

Walker metric as a function of redshift.<br />

The results in the previous chapter were all derived directly from the assumptions of<br />

isotropy and homogeneity. To make further progress we need to determine both R(τ)<br />

and k/A 2 and for this we need to have a theory of gravity and an equation of state to<br />

describe the Universe. The construction, and solution, of Einstein's Field Equations<br />

are beyond the scope of this course. We will therefore simply state the basic<br />

dynamical equation that results from a rigourous treatment in General Relativity and<br />

then seek to justify this in terms of firstly a purely Newtonian analysis and then a<br />

simple-minded approach to General Relativity.<br />

The following equation for R(τ) is the solution to the Field Equations and is called the<br />

Friedmann Equation.<br />

2<br />

(3.1) R & 2 8πG kc R<br />

2 Λ<br />

= ρ − + R<br />

2<br />

3 A 3<br />

2<br />

Hre R(τ) is the scale factor and k/A 2 the curvature introduced in the previous chapter,<br />

ρ is the density of the Universe. The Λ-term is the so-called cosmological constant,<br />

originally omitted by Einstein but then introduced since it enabled him to find a static<br />

solution for the Universe (thereby missing the considerable scientific coup of<br />

predicting the non-static nature of the Universe). We'll see below that it can be viewed<br />

as a curvature of the vacuum. After it was realized that the Universe is not static, it<br />

became conventional to set Λ to zero. However, it has now reappeared in a new and<br />

interesting guise.<br />

3.1 An almost-completely Newtonian approach<br />

Let us first see how far we can get with a purely Newtonian approach. It turns out<br />

that if the Universe is made up of dust (i.e. a pressure-less material), a purely<br />

Newtonian treatment of gravity produces the correct expressions for R(τ), but of<br />

course fails to describe the curvature k/A 2 in the Friedmann equation.<br />

Consider the classical Newtonian approach to calculating the motion of a galaxy A<br />

located a physical distance x from us. Newton's two theorems concerning the


gravitational effects of uniformly distributed distributions of mass allow us to<br />

consider only the mass enclosed within the sphere of radius x = R(τ)ω OA centered on<br />

O whan calculating the gravitational attraction between O and A..<br />

O<br />

x<br />

A<br />

Thus if x is the distance OA, then we have<br />

(3.2) &&x =− 4 π<br />

G ρ x<br />

3<br />

Now, since the comoving distance to A from O, ω OA , is by definition constant during<br />

the expansion, we can write:<br />

x = Rω ⇒ x& = R& ω ⇒ x&& = R&&<br />

ω<br />

OA OA OA<br />

We may thus rewrite our dynamical equation in terms of R(τ) alone, thus eliminating<br />

reference to any particular galaxy or distance x.<br />

(3.3) R && 4π<br />

=− Gρ<br />

R<br />

3<br />

So, the dynamics of our shell of material are the same as the dynamics of the whole<br />

Universe. In a sense, this is then the justification for the Newtonian approach since we<br />

can make our radius r arbitrarily small so that curvature and light travel time effects<br />

are completely irrelevant.<br />

To integrate this equation we need to know how ρ behaves as functions of R(τ), i.e.<br />

we need to know the equation of state of the matter-energy in the Universe. For the<br />

time being lets assume it is simple pressure-less matter (known as “dust”) that evolves<br />

as ρ ∝ R -3 .<br />

We adopt the convention of putting in square brackets quantities that are invariant<br />

under the expansion. If we multiply both sides by & R and set [ρR 3 ] = constant, then we<br />

have


(3.4)<br />

&<br />

&&& 4πG R<br />

RR R<br />

3<br />

=− ρ<br />

2<br />

3 R<br />

R&<br />

2 8πG = R<br />

3 1<br />

ρ + 2ξ<br />

3 R<br />

R&<br />

2 8πG = ρR<br />

2<br />

+ 2 ξ<br />

3<br />

Here, ξ is a constant of integration that, in this Newtonian analysis, corresponds to the<br />

binding energy of the system: If ξ > 0, then we have solutions with<br />

&R 2 → 2ξ as R→ ∞. R(τ) is proportional to τ and the Universe is unbound and<br />

expands forever. If ξ < 0, we get R & = 0 at finite R so the Universe is bound and at<br />

some point the expansion is halted and the Universe recollapses. If ξ = 0, then we<br />

have the asymptotic limit R & → 0 as R→ ∞<br />

Note that this has exactly the same form as the Relativistic Friedmann equation (3.1)<br />

with Λ= 0 if we associate the ξ binding energy with the curvature term.<br />

Let us now return to the equation of state and consider possibilities other than the<br />

familiar one considered above, where we assumed that the density is dominated by<br />

dust with ρ ∝ R -3 .<br />

An obvious possibility is that the density of the Universe is dominated by<br />

electromagnetic radiation, e.g. the cosmic background radiation field. As the<br />

Universe expands, the density of this radiation field will drop as R -4 . We get a factor<br />

of R -3 from the decreasing number density of photons, but an additional factor of R -1<br />

from the fact that all the photons lose energy due to the redshift.<br />

There are other more exotic possibilities. As we’ll see below, we could imagine a<br />

scalar density field that does not change its density as the Universe expands, i.e. ρ =<br />

constant (more than that – in fact we have evidence that this is the case!). We will<br />

refer to this as a false vacuum density.<br />

Other possibilities exist. A few years ago there was some interest in hypothetical<br />

Universes that were dominated by the mass from “cosmic strings” – strange entities<br />

that have a constant density per unit length regardless of how much they are stretched.<br />

These give ρ ∝ R -2 . It is easy to see that putting these other ρ(R) dependencies into<br />

our Newtonian equation (3.3) will not yield the right form of (3.4) upon integration<br />

unless we introduce a small fudge. To see how to proceed, it is useful to consider the<br />

ρ(R) dependence in thermodynamic terms.<br />

We can assume that the expansion is adiabatic (i.e. that there is no inflow or outflow<br />

of heat). The 1st Law of Thermodynamics gives us that the change in the internal<br />

energy (ρV) equals the pdV work done by the system in expanding:<br />

2<br />

(3.5) c d( ρV)<br />

=−pdV<br />

⇒<br />

dρ<br />

=−<br />

( ρ + pc)<br />

2<br />

3<br />

dR<br />

R<br />

If the pressure can be neglected, then


(3.6)<br />

dρ<br />

dR<br />

=−3<br />

⇒ ρ ∝R<br />

−<br />

ρ R<br />

3<br />

This is thus nothing more than the conservation of mass that we had above for the<br />

“dust” case.<br />

If the Universe is dominated by radiation or other relativistic particles then in this case<br />

p = ρc 2 /3, and<br />

(3.7)<br />

dρ<br />

dR<br />

=−3<br />

⇒ ρ ∝R<br />

4<br />

−4<br />

ρ R<br />

3<br />

This is precisely the R -4 dependence that we considered above, arising from the<br />

decrease in photon number density and the loss of energy of photons.<br />

A more exotic case is when we have “material” that exerts a negative pressure, p = −<br />

ρc2. This then gives<br />

d ρ = 0<br />

i.e. the ρ = constant false vacuum case that we mentioned above.<br />

It is trivial to convince yourself that when we we integrate the Newtonian && R equation<br />

(3.3) we will after all get the correct Relativistic Friedmann equation if we introduce a<br />

small fudge: In the Newtonian && R equation, we should use what is called the active<br />

density ρ' defined as follows:<br />

p<br />

(3.8) ρ′ = ρ+<br />

3<br />

c 2<br />

where ρ is the proper density of mass-energy in the Universe and p is the pressure. It<br />

should be stressed that the presence of a pressure term has nothing to do with the<br />

effects of pressure in classical dynamics, since it is pressure gradients that have<br />

dynamcal effects, and there will be no pressure gradients in a homogeneous Universe.<br />

For a radiation field, this has the effect of making this co-called active density twice<br />

the real density. For a false vacuum, the active density becomes negative, so we<br />

would anticipate a repulsive effect for gravity.<br />

It is important to note that we needed to make this fudge of using the active density in<br />

the Newtonian && R equation so as to reproduce the correct Friedmann equation for & R.<br />

When we come to solve the Friedmann equation for different equations of state to get<br />

R(τ), we use the normal density regardless of the pressure, but must, of course, use the<br />

appropriate ρ(R) dependence.


3.2 A poor-man’s General Relativistic treatment<br />

After exploring the Newtonian approach above, in this section we will look at a<br />

simple-minded approach to General Relativity. It must be stressed that the solution of<br />

Einstein's field equations involves the association of the curvature tensor of the metric<br />

and the tensor describing the distribution of mass in the Universe and the tensor<br />

calculus involved is beyond the scope of the course. In this section, we will look at a<br />

highly simplified analysis (taken from Berry's book) in which we will consider two<br />

dimensional surfaces in which the curvature may be represented by a single number.<br />

The point of this is simply to see how dynamical information can come from an<br />

analysis of the curvature of space time.<br />

If we suppress the two angular dimensions, then the alternative form of the<br />

Robertson-Walker metric, which has the Euclidean angular part of the metric (which<br />

makes the suppression of the angular part easier) reduces to:<br />

2<br />

2 2 2 R ( τ)<br />

(3.9) ds = c dτ<br />

−<br />

kr A dr<br />

2 2<br />

1 −<br />

Thus the metric coefficients, g µν, are:<br />

2<br />

(3.10)<br />

g<br />

c<br />

2<br />

0<br />

2<br />

µν<br />

= R ( τ )<br />

0 −<br />

2<br />

1−<br />

kr<br />

A<br />

2<br />

Gauss's classic formula for calculating the curvature of a two-dimensional surface<br />

from the metric coefficients is, in the case where g 12 = g 21 = 0:<br />

(3.11)<br />

1<br />

K =<br />

2g<br />

g<br />

⎪⎧<br />

2<br />

∂ g<br />

⎨−<br />

⎪⎩<br />

∂r<br />

2<br />

∂ g<br />

−<br />

2<br />

∂τ<br />

1<br />

+<br />

2g<br />

⎡∂g11<br />

∂g<br />

22 ⎛ ∂g<br />

⎢ + ⎜<br />

⎢⎣<br />

∂τ ∂τ ⎝ ∂r<br />

2<br />

⎞ ⎤<br />

⎟ ⎥ +<br />

⎠ ⎥⎦<br />

11 22<br />

11<br />

2<br />

11 22<br />

11<br />

2<br />

1<br />

g<br />

22<br />

⎡∂g<br />

⎢<br />

⎢⎣<br />

∂r<br />

11<br />

∂g<br />

∂r<br />

22<br />

⎛ ∂g<br />

22<br />

+ ⎜<br />

⎝ ∂τ<br />

2<br />

⎞<br />

⎟<br />

⎠<br />

⎤⎪⎫<br />

⎥⎬<br />

⎥⎦<br />

⎪⎭<br />

If we evaluate this, we find, since all the derivatives of g 11 vanish:<br />

(3.12) K<br />

R<br />

=− && 2<br />

Rc<br />

Knowing a little about GR, we can plausibly associate this curvature of the space-time<br />

surface, which is a scalar quantity, with the mass density of the Universe, another<br />

scalar, plus possibly a constant representing a curvature of the vacuum when the<br />

density is zero. i.e. we can guess that:<br />

R&&<br />

(3.13) K =− = αρ +<br />

β<br />

2<br />

Rc 2


This may be integrated in the usual way to produce:<br />

(3.14) R & 2 c<br />

2 R<br />

2 R<br />

2 c<br />

2<br />

= 2αρ − β − χ<br />

Here χ will be a constant of integration representing the initial conditions.<br />

You will see that this equation is the Friedmann equation if I associate α with<br />

4πG/3c 2 , β, which we introduced as the curvature of the vacuum, with -Λ/3c 2 , and χ<br />

with kc 2 /A 2 .<br />

Only a full solution of the field equations gives the correct value for the constant χ<br />

that gives, through A, the curvature of the constant τ surface.<br />

3.3 Asymptotic solutions to the Friedmann equation<br />

We have from above (3.3) the Friedmann equation. We have three terms on the right<br />

hand side: (a) a density term; (b) a curvature term and (c) the Λ term.<br />

&R<br />

2<br />

8πG kc Λ<br />

= ρR<br />

− + R<br />

2<br />

3 A 3<br />

2 2<br />

2<br />

It is important to notice that a non-zero Λ term has exactly the same effect as setting ρ<br />

= constant in the first, i.e. density, term. For this reason, a scalar field which produces<br />

a false vacuum density that does not decrease with the expansion is often stilll<br />

referred to as “Λ”, even though I would argue that the physical origin is slightly<br />

different and that calling it Λ is slightly confusing. Whether to regard such a<br />

component as a density component with weird behaviour (false vacuum) or as a nonzero<br />

Λ will introduce a possible source of confusion in what follows. I will try to<br />

consistently denote with a subscript different components of the density.<br />

As an introduction, let us look first at the simplest asymptotic solutions to the<br />

Friedmann equation that arise when the different terms on the right hand side<br />

dominate.<br />

3.3.1 Density term dominates with matter ρ ∝R -3<br />

For a regular matter-dominated Universe with k = 0 and Λ = 0, we can solve the<br />

Friedmann equation as follows


R&<br />

2<br />

3<br />

2<br />

0<br />

8πG<br />

=<br />

3<br />

RR&<br />

=<br />

R<br />

R<br />

R<br />

3<br />

2<br />

=<br />

3<br />

[ ρR<br />

]<br />

R<br />

8πG<br />

3<br />

⎛ 3 ⎞<br />

= ⎜ H<br />

0τ<br />

⎟<br />

⎝ 2 ⎠<br />

3<br />

[ ρR<br />

]<br />

⎛ 8πG<br />

⎜ ρ<br />

0R<br />

⎝ 3<br />

2<br />

3<br />

3<br />

0<br />

⎞<br />

⎟τ<br />

⎠<br />

Thus the Universe expands with R proportional to τ 2/3 . The elapsed time, τ 0 , since the<br />

time when R = 0, is:<br />

2 1<br />

(3.25) τ 0<br />

=<br />

3 H0<br />

3.3.2 Density term dominates with radiation ρ ∝R -4<br />

If the density of the Universe is dominated by radiation, or some other relativistic<br />

species, then with energy density ε, then the basic Friedmann equation will be:<br />

4<br />

(3.37) R & G R<br />

2 8π<br />

ε kc<br />

= −<br />

2 2<br />

3c<br />

R A<br />

We can thus solve the Friedmann equation as before:<br />

2<br />

2<br />

(3.38)<br />

G<br />

RR& 8π<br />

=<br />

2<br />

3c<br />

εR<br />

8πG<br />

4<br />

R = 4 εR<br />

2<br />

3c<br />

τ<br />

8πG<br />

R=<br />

R 4<br />

0 2 ε<br />

0<br />

3c<br />

τ<br />

4<br />

The Universe expands as τ 1/2 , and the age of the Universe is given by<br />

(3.39) τ = 1<br />

2H<br />

3.3.3 Curvature term dominates<br />

The completely undecelerated case with ρ = 0 (and Λ = 0) is trivial to solve:


(3.20)<br />

R&<br />

=<br />

R<br />

R<br />

0<br />

c<br />

A<br />

= H 0τ<br />

with<br />

k<br />

= −1<br />

This is a linear expansion as expected for an undecellerated Universe. The age is<br />

simply the inverse of H.<br />

1<br />

(3.21) τ 0<br />

= H<br />

0<br />

Note that the curvature in this solution, i.e. k = -1 and c/A = dR/dτ, are exactly as we<br />

found in the Milne model in <strong>Chapter</strong> 2.<br />

3.3.4 Λ-term dominates (or false vacuum density with constant ρ)<br />

Let us first consider the non-zero Λ case. From the form of the Friedmann equation,<br />

we will get:<br />

(3.22)<br />

2 1<br />

R&<br />

= ΛR<br />

3<br />

R&<br />

Λ<br />

=<br />

R 3<br />

R<br />

R<br />

0<br />

⎛<br />

= exp⎜<br />

⎝<br />

2<br />

Λ ⎞<br />

( τ −τ<br />

) ⎟<br />

0<br />

3<br />

⎠<br />

Thus, for positive Λ, we get an accelerating, exponential expansion. Hubble's<br />

parameter is, for once, actually a constant, and is given by:<br />

Λ<br />

(3.23) H = ⇔ R∝<br />

3<br />

exp( H τ)<br />

Note that an identical result is obtained by simply inserting an equation of state such<br />

that ρ = constant. If I denote this constant density by ρ Λ then we get:<br />

(3.24)<br />

H<br />

=<br />

8<br />

Λ<br />

πGρ<br />

3<br />

The accelerating effects under gravity of the weird ρ = constant material should not be<br />

surprising when it is recalled that the “active density” in the original Newtonian && R<br />

equation ρ’ = ρ+3p/c 2 is negative, since p = -rc 2 .<br />

3.3.5 The R dependence of the different terms in the Freiedmann equation<br />

It is instructive to see how the relative importance of the different terms of the<br />

Friedmann equation will change as the Universe changes in size through R(τ).


(a)<br />

Matter vs. Radiation in the density term<br />

If the Universe contains both matter and radiation (as ours most certainly<br />

does) then at some earlier time, radiation will come to dominate the density,<br />

even if at later times, the matter dominates. The difference of one power of R<br />

in the behaviour of ρ(R) means that radiation will dominate for R < R eq :<br />

(3.25)<br />

R eq<br />

=<br />

ρ<br />

ρ<br />

rad<br />

matter<br />

0<br />

(b)<br />

Matter vs. curvature<br />

For regular pressure-less dust, the density term may be written [ρR 3 ]/R,<br />

whereas the curvature term is independent of R. Thus as R increases, the<br />

effects of curvature will become more important, as R. Conversely, as we<br />

look back to smaller R, the effects of curvature will become less and less<br />

important. Even if curvature dominates today, at some earlier epoch curvature<br />

will have been negligible, but the relative importance of matter and curvature<br />

will only change as a single power of R.<br />

If the density is dominated by radiation, then the decline of the relative<br />

importance of the curvature term will be more rapid, going as R 2 .<br />

(c)<br />

Matter vs. Λ or false vacuum energy<br />

As with curvature, as we look back to smaller R for a Universe that today<br />

contains both matter (or radiation) and false vacuum energy (or non-zero Λ),<br />

we will find at smaller R that the latter becomes less important. However, the<br />

transition to a matter- (or radiation-) dominated Universe is much faster than<br />

was the case for curvature, going as R 3 (or R 4 for radiation).<br />

3.4 General solutions: the cosmological density parameter Ω and the<br />

deceleration parameter q<br />

It is convenient to introduce a critical density, ρ c and a density parameter Ω, which is<br />

simply the ratio of the actual density to that critical density:<br />

(3.26)<br />

2<br />

3H<br />

ρc<br />

=<br />

8πG<br />

ρ<br />

Ω =<br />

ρ<br />

c<br />

Dividing by R 2 , we can rewrite the Friedmann equation to give a relationship between<br />

the density parameter, the expansion rate and the curvature. Note that the density<br />

here is the sum of all components; matter, radiation, even false vacuum. In what<br />

follows, I will assume that the Universe may contain significant amounts of (pressure-


less) matter (which could be baryonic or non-baryonic), radiation (with pressure) and<br />

false-vacuum (with negative pressure and ρ = constant). I will neglect any other even<br />

more exotic components, including the possibility that the false-vacuum changes with<br />

R. However, it should be clear how to include any other general term in the<br />

discussion below.<br />

Neglecting a Λ-term for the time being (i.e. putting false vacuum, if relevant, into the<br />

density ρ tot to compute an Ω tot ), i..e.<br />

ρ<br />

Ω<br />

tot<br />

tot<br />

= ρ + ρ + ρ<br />

= Ω<br />

m<br />

m<br />

r<br />

+ Ω<br />

r<br />

Λ<br />

+ Ω<br />

Λ<br />

we have<br />

(3.27)<br />

H<br />

2<br />

k<br />

2<br />

R A<br />

( RA)<br />

= H<br />

2<br />

2<br />

=<br />

2<br />

⎛<br />

= ⎜<br />

⎝<br />

Ω<br />

H<br />

c<br />

2<br />

tot<br />

2<br />

c<br />

H<br />

( Ω<br />

2<br />

⎞<br />

⎟<br />

⎠<br />

2<br />

kc<br />

−<br />

2<br />

A R<br />

tot<br />

( Ω<br />

2<br />

−1)<br />

k<br />

−1)<br />

tot<br />

Thus the critical density is also the density of a Universe with zero spatial curvature.<br />

As we’ll see later, (c/H) is a basic length scale in the Universe (it is obviously roughly<br />

c times the age of the Universe τ) and the value of Ω tot tells us about the (physical)<br />

radius of curvature of the Universe, RA, compared with c/H.<br />

Note again that for ρ = 0, i.e. Ω tot = 0, and k = -1, equation (3.16) reduces to A -1 =<br />

HR/c, which is exactly the expression that we derived in our Special Relativistic<br />

analysis.<br />

Note as an aside that if we had chosen to keep the Λ-term in the Friedmann equation<br />

(i.e. to consider the density only to be made up of “normal” matter and radiation),<br />

then the above equations become:<br />

(3.28)<br />

2<br />

2 2 kc Λ<br />

H = H Ω − +<br />

2 2<br />

A R 3<br />

2<br />

k H ⎛ Λ<br />

= ( 1)<br />

2 2 2<br />

⎜ Ω − +<br />

R A c ⎝ 3H<br />

( RA)<br />

2<br />

⎛<br />

= ⎜<br />

⎝<br />

c<br />

H<br />

2<br />

⎞<br />

⎟<br />

⎠<br />

k<br />

( Ω −1+ Λ / 3H<br />

2<br />

⎞<br />

⎟<br />

⎠<br />

2<br />

)<br />

The quantity Λ/3H 2 is often written as λ. The condition for a spatially flat Universe<br />

with k = 0 is thus Ω tot = 1 if I include possible false vacuum energy in my<br />

computation of ρ and hence Ω tot , or Ω + Λ/3H 2 = 1 if I do not and consider instead a<br />

non-zero Λ. These are very important relations, worthy of there own line in the notes.


(3.29) Conditions for a flat Universe:<br />

Ω<br />

tot<br />

= 1<br />

( Ω + λ)<br />

= 1<br />

There is another way to look at Ω. In terms of regular matter and radiation where we<br />

have a decelerating Universe, using the definition of H from the previous chapter we<br />

can rewrite the Friedmann equation in its Newtonian form with the binding energy as:<br />

(3.30) 2 ξ<br />

H<br />

2 (1 − Ω)<br />

=<br />

2<br />

R<br />

Clearly, if ρ < ρ c , i.e. Ω < 1, then ξ > 0 and the Universe is unbound and if ρ > ρ c, i.e .<br />

Ω > 1 then ξ < 0. and the Universe will recollapse at some point in the future. Thus<br />

the critical density is simply the density of a Universe of zero binding energy. The<br />

concept of binding energy with a false vacuum component is not so well defined.<br />

As noted in the previous lecture, we can also define a dimensionless deceleration<br />

parameter, q,<br />

RR<br />

(3.31) q =− &&<br />

R& 2<br />

It is easy to see that the Newtonian R & equation, with the “correct” ρ’ = ρ + 3p/c 2 ,<br />

reduces to:<br />

4π<br />

3<br />

(3.32) R &<br />

= − GR( ρ + 2ρ<br />

− 2ρ<br />

)<br />

so that<br />

Ω<br />

m<br />

(3.30) q = + Ω<br />

r<br />

− Ω<br />

Λ<br />

2<br />

m<br />

r<br />

Remembering that a flat Universe has Ω tot = 1, the condition for a flat Universe<br />

becomes:<br />

3Ω<br />

m<br />

(3.31) q = + 2Ω<br />

r<br />

−1<br />

2<br />

The relationship between Ω and q is a test of the equation of state of the Universe.<br />

Note that Ω 0 is a density measurement that can in principle be determined locally<br />

(questions of homogeneity aside) whereas q 0 is a kinematic measurement that requires<br />

us to observe sufficiently far away that we can measure the second derivative of R(τ).<br />

3.5 General Solutions to the Friedmann equation<br />

Λ


We may write the density of the Universe at any R in terms of the present density (i.e.<br />

at R = R 0 ) which is parameterized by Ω 0 for the different components.<br />

(3.32)<br />

8πGρ<br />

3H<br />

2<br />

0<br />

⎛<br />

⎜Ω<br />

⎜<br />

⎝<br />

⎛ R ⎞<br />

⎜<br />

⎟<br />

⎝ R0<br />

⎠<br />

−3<br />

+ Ω<br />

⎛ R ⎞<br />

⎜<br />

⎟<br />

⎝ R0<br />

⎠<br />

−4<br />

+ Ω<br />

=<br />

0, m<br />

0, r<br />

0, Λ<br />

⎞<br />

⎟<br />

⎟<br />

⎠<br />

We can introduce a new time coordinate, η, called the conformal time though we<br />

won’t encounter it much, such that<br />

(3.33)<br />

cdt<br />

d η =<br />

AR<br />

so that, indicating differentiation w.r.t η with a prime, the Freidmann equation<br />

becomes:<br />

(3.34)<br />

⎛ R′<br />

⎞<br />

⎜<br />

⎟<br />

⎝ R ⎠<br />

2<br />

8πGρ<br />

2<br />

= A R<br />

2<br />

c<br />

0 3<br />

2<br />

⎛<br />

⎜<br />

⎝<br />

R<br />

R<br />

0<br />

⎞<br />

⎟<br />

⎠<br />

2<br />

⎛<br />

− k<br />

⎜<br />

⎝<br />

R<br />

R<br />

0<br />

⎞<br />

⎟<br />

⎠<br />

2<br />

We can now substitute in the expression above for ρ(R)<br />

⎛ R′<br />

⎞<br />

⎜<br />

⎟<br />

⎝ R0<br />

⎠<br />

2<br />

=<br />

H<br />

c<br />

2<br />

0<br />

2<br />

⎛<br />

⎜Ω<br />

⎜<br />

⎝<br />

0, Λ<br />

+ Ω<br />

0, m<br />

⎛<br />

⎜<br />

⎝<br />

R<br />

R<br />

0<br />

⎞<br />

⎟<br />

⎠<br />

−3<br />

+ Ω<br />

0, r<br />

⎛<br />

⎜<br />

⎝<br />

R<br />

R<br />

0<br />

⎞<br />

⎟<br />

⎠<br />

−4<br />

⎞<br />

⎟ 2<br />

A R<br />

⎟<br />

⎠<br />

2<br />

0<br />

⎛<br />

⎜<br />

⎝<br />

R<br />

R<br />

0<br />

⎞<br />

⎟<br />

⎠<br />

4<br />

⎛<br />

− k<br />

⎜<br />

⎝<br />

R<br />

R<br />

0<br />

⎞<br />

⎟<br />

⎠<br />

2<br />

and can clean this up with a = R/R 0 and substitute in the (3.27) relation:<br />

k H<br />

0<br />

(3.35) = ( Ω0,<br />

−1)<br />

2 2 2 tot<br />

R A c<br />

and rearrange to get:<br />

0<br />

2<br />

(3.36)<br />

a′<br />

2<br />

=<br />

k<br />

( ) ( 2<br />

4<br />

Ω<br />

)<br />

0, r<br />

+ Ω0,<br />

ma<br />

− ( Ω0,<br />

tot<br />

−1)<br />

a + Ω0,<br />

Λa<br />

Ω −1<br />

0<br />

This is straightforward to solve provided that Ω 0,Λ = 0 (see 3.6 below).<br />

Taking this equation and substituting in the relation for ρ(R) above (3.32) it is easy to<br />

show that H(R) is given by:<br />

2 2<br />

−3<br />

−4<br />

−2<br />

(3.37) H = H ( Ω + Ω a + Ω a − ( Ω −1)<br />

a )<br />

0<br />

0, λ<br />

0, m<br />

0, r<br />

0, tot<br />

This tells us the value of Hubble’s parameter at all epochs.


Now, since from the Robertson-Walker metric, we have for the path of a photon that<br />

(3.38)<br />

cdR<br />

Rd ω = cdτ<br />

= =<br />

R&<br />

cdR<br />

HR<br />

We can calculate the rate at which a photon travels over comoving radial distance in<br />

terms of the observable redshift, by noting that R=R 0 /(1+z). Thus<br />

c<br />

H<br />

c<br />

H<br />

2<br />

3<br />

4<br />

(3.39) R dω<br />

= dz = [(1<br />

− Ω )(1 + z)<br />

+ Ω + Ω (1 + z)<br />

+ Ω (1 z ] dz<br />

0 0, tot<br />

0, Λ 0, m<br />

0, r<br />

+ )<br />

0<br />

We would now in principle be able to calculate ω(z) and, knowing the curvature term<br />

S k (ω) and k/A 2 , compute things like D(z), D L (z), D θ (z), and the incremental volume<br />

element dV/dz, etc. These are all quantities that we need to know to make sense of the<br />

distant Universe and/or test the cosmology or determine the cosmological parameters<br />

from observation. Unfortunately, simple analytic solutions are only derivable for the<br />

case of Ω 0,Λ = 0.<br />

Finally, since we always have<br />

2<br />

kc<br />

(3.40) Ω<br />

tot<br />

−1<br />

=<br />

2 2 2<br />

H R A<br />

we can calculate Ω tot at all R.<br />

−0.5<br />

(3.41)<br />

Ω<br />

tot<br />

−1<br />

=<br />

−2<br />

2<br />

( 1− Ω + Ω (1 + z)<br />

+ Ω (1 + z)<br />

+ Ω (1 + z)<br />

)<br />

0, tot<br />

0, Λ<br />

Ω<br />

0, tot<br />

−1<br />

0, m<br />

0, r<br />

An important point is that unless Ω 0,m ~ Ω 0,r ~ 0 (i.e. unless we are completely<br />

dominated by a false vacuum at the present epoch) then at very high redshifts, all<br />

Universes will have Ω tot ~ 1, regardless of their current Ω 0,m , Ω 0,r or Ω 0,Λ .<br />

3.6 Case study: matter dominated Universes<br />

The treatment of a matter dominated Universe of arbitrary density is relatively simple<br />

analytically, allowing us to explore several topics addressed in the general case in the<br />

previous section quite easily. Until very recently, we were fairly confident that this<br />

also in fact described our own Universe. Within the last 2-3 years, there has emerged<br />

evidence, that certainly deserves to be taken seriously, that there is a dominant false<br />

vacuum term (or non-zero Λ). If correct, this makes the general solution for R(τ) and<br />

k/A 2 much messier as discussed above. However, it is still worthwhile looking at the<br />

matter-dominated case.<br />

The rest of this extensive section deals with such a Universe (which may be<br />

hypothetical). In this section Ω and Ω 0 will be taken be taken to be the current density<br />

of matter, i.e. what we called Ω m above.


3.6.1 General solutions for R(τ)<br />

Before looking at general solutions to the Friedmann equation, it is useful to look at<br />

the asymptotic behaviour of models with general Ω. Setting Λ = 0, we may develop<br />

the Friedmann equation, using the relations between curvature and Ω (equation 3.27)<br />

and between R and z, as follows:<br />

(3.42)<br />

Thus<br />

R&<br />

R&<br />

R<br />

2<br />

2<br />

2<br />

0<br />

8π<br />

GρR<br />

kc<br />

= −<br />

3 R A<br />

3 2<br />

8π<br />

Gρ0R0<br />

= −( Ω −1)<br />

H<br />

3 R<br />

2 R0<br />

= H0<br />

Ω0<br />

−( Ω −1)<br />

H<br />

R<br />

2<br />

= H Ω ( 1+ z)<br />

− Ω + 1<br />

0<br />

2<br />

0 0<br />

2<br />

0 0<br />

2<br />

0 0<br />

(3.43)<br />

R&<br />

R<br />

2<br />

2<br />

= H ( 1 +Ω z)<br />

2 0<br />

0<br />

0<br />

Hence, if Ω 0 < 1, we can see that the Universe has had roughly constant & R (i.e. has<br />

had undecelerated expansion) since the epoch corresponding to (1+z) ~ Ω 0<br />

-1. At<br />

earlier times, & R behaves as for an Ω = 1 Universe (though, note, with a different H<br />

than the Ω 0 = 1 Universe that would have the same H 0 today).<br />

We can see this also by looking at Ω(z), which may be calculated as follows: From<br />

(3.43) we have<br />

2 2 2<br />

H = H ( 1+ z) ( 1+<br />

Ω z)<br />

0<br />

0<br />

Substituting this expression for H(z) into the definition of Ω and knowing ρ(R), we<br />

get<br />

Ω =<br />

=<br />

8πGρ<br />

2 2<br />

3H<br />

( 1+ z) ( 1+<br />

Ω z)<br />

0<br />

8πGρ<br />

( 1+<br />

z)<br />

0<br />

2<br />

3H<br />

( 1+<br />

Ω z)<br />

0<br />

0<br />

0<br />

Thus<br />

(3.44) Ω Ω ( 1 + z)<br />

=<br />

0<br />

( 1 + Ω z)<br />

0


Note that if the density was exactly critical at some point then it will be critical for all<br />

later time (as is obvious from the identification of the critical density as that density<br />

which produces the critically expanding Universe). Of more interest, if Ω 0 < 1, then<br />

Ω was approximately unity for all z > Ω 0<br />

-1 as we discussed above.<br />

Thus, an Ω 0 < 1 Universe behaves to first order like a critical flat Universe for (1+z) ><br />

Ω 0<br />

-1 and then like an undecelerated Universe for (1+z) < Ω 0<br />

-1. This approximation<br />

can simplify many calculations, particularly those involving the growth of density<br />

inhomogeneities in the Universe since the growth modes in Ω ~ 1 Universes and<br />

those in Ω ~ 0 Universes are particularly easy to calculate.<br />

One final interesting result is that, if Ω 0 > 1, then the radius of maximum expansion,<br />

i.e. the point at which Η = 0, is given by the condition:<br />

(3.45)<br />

1+ Ω z = 0<br />

R<br />

max<br />

R<br />

0<br />

0<br />

max<br />

Ω0<br />

=<br />

Ω −1<br />

0<br />

We gave above the solution for matter-dominated Ω = 1 (section 3.31). In the case of<br />

a general Ω, parametric solutions may be obtained for R(τ).<br />

It helps to define "natural" units for R and τ as follows in terms of a quantity M<br />

τ 3<br />

η =<br />

2GM c<br />

R 2<br />

ζ =<br />

2GM Ac<br />

4π M = ρ R A<br />

3<br />

3 3<br />

Then the basic Friedmann equation (3.1) becomes<br />

2<br />

c ⎛ dζ<br />

⎞<br />

⎜ ⎟<br />

2<br />

A ⎝ dη<br />

⎠<br />

⎛ dζ<br />

⎞<br />

⎜ ⎟<br />

⎝ dη<br />

⎠<br />

2<br />

2<br />

1 c<br />

=<br />

ζ A<br />

1<br />

= − k<br />

ζ<br />

ζ dζ<br />

= dη<br />

1−<br />

kζ<br />

2<br />

2<br />

kc<br />

−<br />

A<br />

We now introduce a secondary variable, u, such that<br />

2<br />

ζ = S ( u) ⇒ dζ<br />

= 2S ( u) C ( u)<br />

du<br />

k<br />

2<br />

2<br />

k<br />

k


Here S k (u) and C k (u) are analogous to S k (ω) and are sin, sinh, cos, cosh etc.<br />

depending on the value of k. Then<br />

2<br />

Sk<br />

( u)<br />

C<br />

dη<br />

= 2<br />

1−<br />

kS<br />

= 2S<br />

2<br />

k<br />

( u)<br />

du<br />

= k(1<br />

− C<br />

k<br />

⎛ S<br />

η = k⎜u<br />

−<br />

⎝<br />

2<br />

k<br />

( u)<br />

du<br />

( u)<br />

(2u))<br />

du<br />

k<br />

k<br />

(2u)<br />

⎞<br />

⎟<br />

2 ⎠<br />

Now making a final substitution of Θ = 2u, we get parametric solutions<br />

k<br />

η = ( Θ−Sk<br />

( Θ))<br />

2<br />

k<br />

ζ = ( −Ck<br />

2 1 ( Θ))<br />

So, we finally have for R and τ:<br />

(3.46)<br />

kGM<br />

R =<br />

2<br />

Ac<br />

kGM<br />

τ =<br />

3<br />

c<br />

{ 1−<br />

C ( Θ)<br />

}<br />

k<br />

{ Θ − S ( Θ)<br />

}<br />

k<br />

Note that dτ = AR/c dΘ. The parameter Θ is (as noted above) called the "conformal<br />

time" (although we will not make great use of it), and dω is proportional to dΘ for a<br />

light ray.<br />

3.6.2 Determining S k (ω) in the matter-dominated case<br />

We saw in the previous chapter how the effective distance D enters into almost all the<br />

relations between the intrinsic and observed properties of objects. Recall that D = R 0<br />

S k (ω). Thus we need to know S k (ω) or rather S k (z) since the redshift z is the only real<br />

observable related to distance.<br />

We want to get ω(z), which we can get by integrating dω/dz, which in turn can be<br />

obtained by recognizing that:<br />

dω<br />

dz<br />

dω<br />

dt<br />

= × ×<br />

dt dR<br />

We had from equation (3.21):<br />

dR<br />

dz<br />

(3.47)<br />

R&<br />

R<br />

2<br />

2<br />

= H ( 1+Ω<br />

z)<br />

2 0<br />

0<br />

0


Now, using the definition of z,<br />

R0 R0<br />

R = ⇒ dR = −<br />

( 1+<br />

z) ( 1+<br />

z)<br />

2<br />

dz<br />

We get:<br />

(3.48)<br />

dz<br />

H z z<br />

dτ =− + 2<br />

+<br />

0<br />

( 1 ) ( 1 Ω<br />

0<br />

This equation may be integrated to yield τ(z), the age of the Universe as a function of<br />

z and Ω 0 .<br />

Since light propagates along null-geodesics with ds = 0, we have from the metric that<br />

light following a radial trajectory towards us has:<br />

d c<br />

(3.49) Rdω<br />

=− cdτ<br />

⇒ ω<br />

dτ<br />

=− R ( τ )<br />

Thus we can get the change in comoving radial distance along a light ray with<br />

redshift, z,<br />

(3.50) R d ω c 1<br />

0 =−<br />

dz H ( 1+ z) 1+<br />

Ω z<br />

0 0<br />

This expression can be integrated to give ω(z). Then, since we know the functional<br />

form of S k (ω) (= A sin ω/A or whatever, see 2.7) and we know A in terms of Ω 0 or q 0 ,<br />

we can find S k (ω) in terms of the observable z and q 0 .<br />

The quantity R o S k (ω) which we identified in <strong>Chapter</strong> 2 as the effective distance D, is<br />

often written in terms of a function Z q (z) as follows:<br />

c<br />

(3.51) D= R S<br />

H Z z<br />

0 k<br />

( ω ) =<br />

q<br />

( )<br />

0<br />

Recall that at very low redshifts, d = cz/H o , so Z q (z), a function of q 0 , is a kind of<br />

effective redshift which makes the expressions for D look familiar.<br />

For the k = 0 Euclidean case ( Ω = 1), we can integrate (3.32) to give ω(z), and hence<br />

since S k (ω) = ω, we have<br />

(3.52)<br />

0<br />

2 −1/<br />

2<br />

c ⎡ 1 ⎤<br />

D = R0ω<br />

= ⎢ = 2(1 − (1 + )<br />

1/ 2 ⎥<br />

z<br />

H<br />

0 ⎣ ( 1+<br />

z)<br />

⎦<br />

z<br />

)<br />

The actual derivation of Z q (z) in the general case is tedious and extremely messy.<br />

However the final result comes out as:


1<br />

(3.53) z)<br />

= { q z + ( q −1)(<br />

1+<br />

2q<br />

z −1}<br />

Z q<br />

(<br />

2<br />

0 0<br />

0<br />

q0<br />

(1 + z)<br />

This reduces in three simple cases to:<br />

q<br />

0<br />

= 0<br />

k<br />

= −1<br />

Z<br />

q<br />

z(1<br />

+ 0.5z)<br />

( z)<br />

=<br />

(1 + z)<br />

(3.54)<br />

q<br />

q<br />

0<br />

0<br />

= 0.5<br />

= 1<br />

k = 0<br />

k = 1<br />

⎛<br />

Z<br />

q<br />

( z)<br />

= 2⎜1<br />

−<br />

⎝<br />

z<br />

Z<br />

q<br />

( z)<br />

=<br />

1+<br />

z<br />

1 ⎞<br />

⎟<br />

1+<br />

z ⎠<br />

These expressions are very useful. The q 0 = 1 case is noteworthy primarily because it<br />

produces a luminosity distance D L (see section 2.26) that is proportional to z. Thus the<br />

magnitude-redshift relation (the classical Hubble diagram) is a straight line in this<br />

cosmology.<br />

It should be stressed that all of the foregoing relations apply only to a pressureless<br />

matter-dominated Universe, like ours at the present epoch, because they were based<br />

on the particular form of R(t) that is produced in such a model.<br />

The various relations above are sufficient to derive a number of interesting quantities.<br />

For instance, one often encounters the comoving volume element, dV c /dz. This<br />

describes the incremental increase in comoving volume (i.e. in which galaxies have<br />

constant number density assuming that their numbers are conserved) with redshift and<br />

is required for instance when calculating the expected number of faint galaxies seen<br />

within a survey of a given surface area on the sky since dN/dz is proportional to<br />

dV c /dz.<br />

Consider a cone of solid angle dΞ. If this projects to a physical area A on a sphere of<br />

constant radius at a redshift z, then<br />

(3.55)<br />

dV<br />

dz<br />

dV<br />

dz<br />

c<br />

c<br />

2 dw<br />

2<br />

= A( 1+<br />

z)<br />

with A = dΞDθ<br />

dz<br />

⎛<br />

=<br />

⎜<br />

⎝<br />

c<br />

H<br />

0<br />

⎞<br />

⎟<br />

⎠<br />

3<br />

Z<br />

(1 + z)<br />

2<br />

q<br />

( z)<br />

dΞ<br />

1+ Ω z<br />

0<br />

3.7 The interrelation between the curvature, the density and the expansion<br />

rate<br />

Remember that all solutions to the Friedmann equation must have the same<br />

relationship (3.17) between the curvature and the expansion rate and the density:


k<br />

( RA)<br />

2<br />

( Ω −1)<br />

=<br />

2<br />

⎛ c ⎞<br />

⎜ ⎟<br />

⎝ H ⎠<br />

with<br />

Ω= 8 πG<br />

ρ<br />

3H<br />

2<br />

Here ρ is the density, be it that of normal matter density, radiation density or false<br />

vacuum energy density, or the sum of all three if appropriate. Indeed, in the general<br />

case we have:<br />

k<br />

( RA)<br />

2<br />

( Ωtot<br />

=<br />

⎛ c<br />

⎜<br />

⎝ H<br />

−1)<br />

2<br />

⎞<br />

⎟<br />

⎠<br />

with<br />

Ω<br />

tot<br />

8πG<br />

=<br />

2<br />

3H<br />

∑<br />

i<br />

ρ<br />

i<br />

This inter-relationship, which can be written in several different ways, including that<br />

of the original Friedmann equation (3.1), is the heart of the General Relativistic<br />

Friedmann-type models.<br />

Having been brought up since the cradle on Newtonian ideas, we tend to think of<br />

densities and velocities as the key quantities to focus on, with gravity acting on the<br />

density to decellerate the Universe and so on. We think intuitively of Newtonian<br />

concepts such as escape speeds, binding energies, etc. For most of us, the curvature<br />

is then added on as an uncomfortable afterthought. This Newtonian approach is a<br />

useful and pragmatic way to approach questions such as the appearance of distant<br />

objects and the physical evolution of the contents of the Universe. Indeed, as we saw,<br />

we can derive perfectly correct expressions describing the expansion of R(τ).<br />

However, in a fundamental way it is really the other way around. The one quantity<br />

that can never change during the expansion of a homogeneous Universe is the<br />

comoving curvature k/A 2 . It is the curvature that describes the Universe. Once the<br />

curvature is defined, the density at any epoch follows from the expansion rate and<br />

vice versa from (3.17). As the Universe expands, the equation of state tells us how the<br />

density will change (e.g. as R -3 , R -4 and R 0 in the three cases above) and, as if by<br />

magic (but not really of course), the expansion rate is decelerated or accelerated in the<br />

ways that we calculated above to compensate.<br />

In the evolution of the Universe, the effective equation of state (i.e. that of the<br />

dominant density component) has changed and may change again. Each of these<br />

changes brings a different form of R(τ) and of ρ(R) with smooth transitions in each so<br />

as to preserve the correct relationship between them.<br />

However, the topology of the Universe, represented in the homogeneous case by the<br />

comoving curvature scalar k/A 2 , remains constant.


3.8 The Big Bang<br />

We have now calculated R(τ) for three generic models for a pressure-less expanding<br />

Universe. Because gravity acts on matter to decelerate it, it should not be surprising<br />

that these three all have R = 0 at some finite time in the past. As noted above, any<br />

curvature present today will be less important at earlier epochs, as Ω → 1. We would<br />

expect radiation to become dynamically dominant at some earlier point in time, but it<br />

too has R = 0 at some point in the past. Of course, R = 0 implies a singularity and it is<br />

likely that some new physics must be introduced at such early times (as we’ll see<br />

below) but a compact, rapidly expanding state for the Universe is a strong prediction<br />

of the Friedmann models.<br />

As noted above, even a non-zero false vacuum density today would, if constant, be<br />

unimportant before some earlier epoch.<br />

The idea that the Universe was once in an extremely compressed state is the<br />

fundamental feature of the Friedmann models. This compressed but rapidly expanding<br />

state is known as the Big Bang. The term was introduced, in a derisory way, by Fred<br />

Hoyle during a radio broadcast.<br />

<strong>Chapter</strong> 3: Key points<br />

1. The correct Friedmann equation in & R (from GR) can in fact<br />

be generated from the Newtonian && R equation if we set the density to<br />

be the so-called "active density".<br />

2. The dynamics of the Universe depend on the equation of state<br />

of the matter-radiation in the Universe as this determines how the<br />

density changes as the Universe expands. Our matter-dominated<br />

Universe was once radiation dominated.<br />

3. The density parameter Ω determines in fundamnetal way<br />

both the dynamics and the curvature of the Universe.<br />

4. For a given equation of state, the term in the Robertson-<br />

Walker metric mat be written in terms of (c/H 0 )Z q (z), where q is a<br />

deceleration parameter.<br />

5. A false vacuum energy density is equivalent to a non-zero Λ<br />

term in the Friedmann equation.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!