0% found this document useful (0 votes)
2K views59 pages

Ryden Solutions

Uploaded by

iantyler329
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
2K views59 pages

Ryden Solutions

Uploaded by

iantyler329
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 59

Ryden Solutions

Rio Weil
This document was typeset on May 3, 2022

Solutions to problems from Ryden’s “Introduction to Cosmology”, 2nd ed.

Contents
1 Introduction 3
Cosmology FAQ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

2 Fundamental Observations 4
2.1 Human blackbody radiation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.2 CMB photon rate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.3 How long for the CMB to warm you up? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.4 Tired Light . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.5 CMB Cutoff - The Cosmic Infrared background . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.6 CMB Cutoff - The Other Direction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

3 Newton versus Einstein 8


3.1 Evidence for electrical neutrality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
3.2 Angular Width on a sphere . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
3.3 The Earth isn’t flat! . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
3.4 Area Bounds for Equilateral Triangles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
3.5 Equivalent Metrics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

4 Cosmic Dynamics 13
4.1 Does the Cosmological Constant Affect Planetary Motion? . . . . . . . . . . . . . . . . . . . . 13
4.2 Perturbing Einstein’s Static Universe . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
4.3 How Large is Einstein’s Static Universe? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
4.4 Baseballs and Critical Density . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
4.5 Equation of State for Gases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
A Force-Based Derivation of the Newtonian Acceleration Equation . . . . . . . . . . . . . . . 16

5 Model Universes 17
5.1 Redshift in single-component universes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
5.2 Redshift Change Timescale for Flat Matter-Only Universe . . . . . . . . . . . . . . . . . . . . 18
5.3 Present age of universe for positively curved matter-only universe . . . . . . . . . . . . . . . 18
5.4 Present age of universe for negatively curved matter-only universe . . . . . . . . . . . . . . . 20
5.5 Phantom Energy and the Big Rip . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
5.6 Pulling an Einstein . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
5.7 Big Crunch . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
5.8 Big Bounce . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
5.9 Ωm,0 for t = H0−1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
5.10 Amounts in the Benchmark model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
Time and Scale Factor for Matter-Only Universe . . . . . . . . . . . . . . . . . . . . . . . . . . 26

1
6 Measuring Cosmological Parameters 28
6.1 Magnitudes and Polar Bear Feet . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
6.2 Angular Size and Polar Bear Feet . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
6.3 Maximizing d A in a flat, single-component universe . . . . . . . . . . . . . . . . . . . . . . . . 29
6.4 Difference between relative and absolute magnitude . . . . . . . . . . . . . . . . . . . . . . . 30
6.5 Surface Brightness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
6.6 Quasar Light Flux . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
6.7 Proper Area of a sphere . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
6.8 Total Intensity of Standard Candles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
6.9 Expansion Switch . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

7 Dark Matter 36
7.1 Dark Matter Candidates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
7.2 Draco Galaxy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
7.3 Gravitational Lensing of Earth . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
7.4 Halo Mass Density . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
7.5 Cluster Collisions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
7.6 Solar vs. CMB neutrinos . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

8 The Cosmic Microwave Background 40


8.1 Baryon-to-Photon Ratio and Recombination Temperature . . . . . . . . . . . . . . . . . . . . . 40
8.2 An ionizing photon per baryon . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
8.3 Completely Helium at Recombination . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
8.4 Distances to Last Scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42

9 Nucleosynthesis and the Early Universe 43


9.1 Mass fraction of Helium with faster decay . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
9.2 Mass fraction of Helium with different rest energies . . . . . . . . . . . . . . . . . . . . . . . . 43
9.3 Helium Increase in Our Galaxy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
9.4 Maximum value for primordial Helium Fraction . . . . . . . . . . . . . . . . . . . . . . . . . . 44
9.5 Neutrino Detection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
Optical Depth of Reionized Material (9.5 in first ed.) . . . . . . . . . . . . . . . . . . . . . . . 45

10 Inflation and the Very Early Universe 47


10.1 Upper limit on Primordial Density . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
10.2 Solving the Monopole Problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
10.3 False Vacuum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
Gamow’s CMB Prediction (10.3 in first ed.) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

11 Structure Formation: Gravitational Instability 52


11.1 Density Fluctuations in a Flat, Matter-Dominated, Contracting Universe . . . . . . . . . . . . 52
11.2 Density Fluctuations in a Nearly Empty, Negatively Curved, Expanding Universe . . . . . . 52
11.3 Photon-Baryon Fluid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
11.4 Milky Way Gravitational Collapse . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
11.5 Coma Cluster Gravitational Collapse . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
11.6 Mean Square Mass Fluctuation and Standard Deviation of Density Field . . . . . . . . . . . . 55

12 Structure Formation: Baryons and Photons 57


12.1 Galaxies more luminous than L . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
12.2 Total Luminosity Density . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
12.3 Do structures of 1017 M exist today? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
12.4 Ripping Apart Galaxies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

2
1 Introduction
Cosmology FAQ
There are no problems in the Introduction chapter in Ryden. However, this seems like the optimal place
to address a bunch of commonly-asked questions in Cosmology.
(a) Where is the center of the Universe?
Answer: Everywhere; the universe is homogenous on large scales, so there is no preferred location as
a center!
(b) Why is the Universe expanding?
Answer: Because it was expanding yesterday (and it was expanding yesterday due to inflation).

(c) What is the Universe expanding into?


Answer: Itself; there is no boundary to the Universe.
(d) Why isn’t the Solar System expanding?
Answer: It actually was in the past, but it has stopped now.

(e) Is the Universe rotating?


Answer: There are modified FRW metrics that allow for special axes (note this breaks the cosmological
principle; if the Universe is described by these metrics, it would be homogenous but not isotropic).
However, a rotating universe would generate spiral patterns on the CMB; the fact that we do not
observe such patterns allows us to constrain global rotation.

(f) What happened before the Big Bang?


Answer: We don’t know! There are theoretical models (such as cyclic Universes), but to a degree this
is an unanswerable question.
(g) How come the CMB photons haven’t outrun the galaxies since the Big Bang?
Answer: The Big Bang doesn’t work that way; the Universe did not start at one point and then expand.
It happened everywhere at once! CMB photons are emitted isotropically, from everywhere; so at every
point in time we receive CMB photons that were further away.
(h) If the Universe is only 14 billion years old, how come we can see objects that are 40 billion light years
away?
Answer: Due to the expansion of the universe, the horizon of the observable universe is not ct0 but
rather ∼ 3ct0 . You can do the integrals to get the exact prefactor.
(i) What is dark matter?
Answer: A substance that makes up 29% of our universe and only interacts through the gravitational
force. We have a better picture of what it isn’t than what it is.

(j) What will happen in the far future of the Universe?


Answer: We don’t know! But in our best current model (the Benchmark model), a heat death is
predicted.

3
2 Fundamental Observations
2.1 Human blackbody radiation
The energy density of blackbody radiation is given by:

eγ = αT 4 (2.1)

Approximating a human being as a sphere with volume 1m3 , an considering that photons travel at speed
c we have that the rate of energy radiation is:

E = cAeγ = cAαT 4 ∼ 2.10 × 103 W (2.2)

2.2 CMB photon rate


We can calculate the photon number density of the CMB to be:

nγ = βT 3 = (2.029 × 107 m−3 K−3 )(2.8255K)3 = 4.107 × 108 m−3 (2.3)

Approximating our body to be a perfect sphere with cross sectional area A ∼ 1m, since photons travel at
a rate cm s−1 the rate at which they pass through us is:

r ∼ cAnγ = (3.00 × 10m s−1 )(1m2 )(4.107 × 108 m−3 ) = 1.23 × 1017 s−1 (2.4)

2.3 How long for the CMB to warm you up?


The energy per photon of the CMB is:

= 1.02 × 10−22 J (2.5)

The energy required to raise my temperature by 1nK is:

∆E = Cm∆T = (4200J kg−1 K−1 )(50kg)(10−9 K) = 2.1 × 10−4 J (2.6)

So solving for the time to heat up by 1 nanoKelvin:

∆E
t1nK = e ∼ 16.8s (2.7)
r nγγ

2.4 Tired Light


We start with the energy loss per unit distance propsed by the “tired light hypothesis”:

dE
= −kE (2.8)
dr
This is the all-too-famous exponential decay ODE, which has solution:

E(r ) = C exp(−kr ) (2.9)

In principle we could use separation of variables to solve the ODE, but let’s just verify that this is the
correct solution:
d 
C exp(−kr ) = −kC exp(−kr ) = −kE X (2.10)
dr

4
Letting E(0) = E0 be the distance of the energy of the photon when emitted (before it has travelled and
lost energy), we get: 
E0 = E(0) = C exp −k(0) = C =⇒ C = E0 (2.11)
Therefore:
E(r ) = E0 exp(−kr ) (2.12)
Now, using the energy-momentum relation and the DeBroglie wavelength relation, we have:

hc
E = pc = (2.13)
λ
So substituting this into (2.12) we get:
hc hc
= exp(−kr ) (2.14)
λr λ0
Where λr is the observed/measured photon wavelength at distance r from the source and λ0 is the photon
wavelength measured at the source. Rearranging we obtain:

λr
= exp(kr ) (2.15)
λ0

Now we recall the definition of redshift:


λr − λ0 λr
z= = −1 (2.16)
λ0 λ0

Substituting this into (2.15) we get:


z = exp(kr ) − 1 (2.17)
Which adding one and taking the logarithm of both sides we get:

log(1 + z) = kr (2.18)

If z  1, then by Taylor expanding to first order we obtain:

log(1 + z) ≈ z (2.19)

So in this limit we have:


z = kr (2.20)
From which we see a linear distance-redshift relation. For the last part of the question, we recall Hubble’s
Law:
H0
z= r (2.21)
c
So comparing this with (2.20), the value of k to yield a Hubble constant of H0 = 68km s−1 Mpc−1 must be:

H0
k= = 2.3 × 10−4 Mpc−1 (2.22)
c

2.5 CMB Cutoff - The Cosmic Infrared background


We start with the number density for CMB photons:

e( f )d f 8π f 2d f
n( f )d f = = 3  (2.23)
hf c exp h f /kT − 1

5
  
Since h f > E0  kT, we have that exp h f /kT  1 and so exp h f /kT − 1 ∼ exp h f /kT . This yields:
 
8π 2 hf
n( f )d f ≈ 3 f exp − df (2.24)
c kT
Now, to get n(h f > E0 ) we integrate this expression from E0 /h to ∞.
Z ∞  
8π 2 hf
n(h f > E0 ) = f exp − df (2.25)
E0 /h c3 kT

Integrating by parts (with u = f 2 , dv = exp −h f /kT ), we have:




∞ Z ∞  !
2 − kT −kT
  
8π hf hf
n(h f > E0 ) = 3 f exp − − 2f exp − df (2.26)
c h kT E0 /h E0 /h h kT

The term at infinity goes to zero, so:


 !
kTE02 2kT ∞
  
8π E0 hf
Z
n(h f > E0 ) = 3 exp − + f exp − df (2.27)
c h3 kT h E0 /h kT
 
hf
We now integrate by parts again (with u = f , dv = exp − kT ) to get:
 
∞ Z ∞  !
8π  kTE02 −kT −kT
    
E0 2kT hf hf
n(h f > E0 ) = 3 exp − + f exp − − exp − df 
c h3 kT h h kT E0 /h E0 /h h kT
(2.28)
Again the term at infinity goes to zero and we get:
 
2     Z ∞   !
8π kTE E0 2kT kTE0 E0 kT hf
n(h f > E0 ) = 3  3 0 exp − + exp − + exp − df  (2.29)
c h kT h h2 kT h E0 /h kT

Finally the last integral is easy:


  
!

8π  kTE02 −kT
     
E0 2kT  kTE0 E0 kT hf
n(h f > E0 ) = exp − + exp − + exp − 
c3
 3 2
h kT h h kT h h kT

E0 /h

(2.30)
So after this tedious calculation, we have:
" #
kTE02 2k2 T 2 E0 2k3 T 3

8π E0
n(h f > E0 ) = 3 exp − + + (2.31)
c kT h3 h3 h3

Using that E0  kT again, we can neglect all but the first term (so we could have really avoided the latter
two integration steps, but alas):

8πkTE02
 
E0
n(h f > E0 ) ≈ exp − (2.32)
c3 h3 kT
Now taking the ratio of this with nγ :

8πkTE02
 
E0
n(h f > E0 ) c3 h3
exp − kT 
E0
2 
E0

≈ 2.4041 k3
= 0.42 exp − (2.33)
nγ T3 kT kT
π 2 h̄3 c3

6
which was the desired formula. Next, we calculate the fraction of “CMB” photons that are actually far-IR
photons. We first consider the wavelength-frequency relation for light:

c = λf (2.34)

so λ < 1mm corresponds to f > 3 × 1011 Hz and so:

E = h f > 2 × 10−22 J (2.35)

So using the above derived relation with T = 2.7255K we have:


!2 !
n(h f > 2 × 10−22 J) 2 × 10−22 J 2 × 10−22 J
≈ 0.42 exp − = 0.058 (2.36)
nγ k · 2.7255K k · 2.7255K

2.6 CMB Cutoff - The Other Direction


We again recall the number density of the CMB photons:

8π f 2d f
n( f )d f = (2.37)
c3 exp h f /kT − 1


Since kT  E0 > h f , we have that:  


hf hf
exp ≈ 1+ (2.38)
kT kT
by Taylor expanding to first order. The number density in this regime therefore becomes:

8π f 2d f 8πkT
n( f )d f ≈ = fdf (2.39)
c3 1 + h f − 1 hc3
kT

Integrating this from 0 to E0 /h, we obtain n(h f < E0 ):

4πkTE02
Z E0 /h E0 /h
8πkT 4πkT 2
n(h f < E0 ) ≈ fdf = f = (2.40)
0 hc3 hc3 0 h3 c3
So solving for the fraction of photons in the CMB with h f < E0 we have:

4πkTE02  2
n(h f < E0 ) h3 c3 E0
≈ 2.4041 k3
= 0.21 (2.41)
nγ T3 kT
π 2 h̄3 c3

Solving for the fraction of CMB photons with λ > 3cm (and hence capable of passing through the Earth’s
atmosphere) we again use the wavelength-frequency relation for light of c = λ f . λ > 3cm corresponds to
f < 1010 Hz so:
E = h f < 6.63 × 10−24 J (2.42)
So using our obtained relation with T = 2.7255K we have:
!2
n(h f < 7 × 10−24 J) 7 × 10−24 J
≈ 0.21 = 0.0065 (2.43)
nγ k · 2.7255K

7
3 Newton versus Einstein
3.1 Evidence for electrical neutrality
Through astronomical observations, we notice that gravitational forces dominate dynamics in the universe
on large scales. However, electrostatic forces are much stronger than gravitational. A simple demonstra-
tion of this is given by comparing the magnitudes of the gravitational and electrostatic forces between
an electron and proton. At a distance r, the gravitational force between an electron and proton (in the
Newtonian picture) is given by:
Gm p me 1 1.01 × 10−66
Fg = 2
= 2 (6.67 × 10−11 N kg−2 m2 )(1.67 × 10−27 kg)(9.11 × 10−31 kg) = N (3.1)
r r r2
And the elecctrostatic force is given by:
ke2 1 2.30 × 10−28
| Fe | = 2
= 2 (8.99 × 109 N C−2 m2 )(1.60 × 10−19 C )2 = N (3.2)
r r r2
Taking the ratio we have:
2.30×10−28
Fe r2
N
= 1.01×10−66
= 2.28 × 1038 (3.3)
Fg N
r2
As we can see from this example, electrostatic forces are much stronger than gravitational (38 orders of
magnitude stronger in this case!). Hence if there was a large charge imbalance in the universe, we would
observe that electrostatic forces would dominate large-scale dynamics rather than gravitational. This
however disagrees with our observations, and we conclude that the universe must be (mostly) electrically
neutral.

3.2 Angular Width on a sphere


We recall that the metric for a sphere of radius R is given by:
 
r
dl 2 = dr2 + R2 sin2 dθ 2 (3.4)
R
So we can write the width dl of the object using the above equation. Since dl  R, we can take the entire
width of the object to be located at the same distance r away from us (the observer). In other words, in
this limit we have dr = 0 and hence the above reduces to:
 
2 2 2 r
dl = R sin dθ 2 (3.5)
R
Isolating for the angular width of the object, we obtain:

dl
dθ = r
 (3.6)
R sin R

As we take r → πR (i.e. the object is located at the antipode) we have:


dl
lim dθ = lim  =∞ (3.7)
r → Rπ R sin Rr
r → Rπ
 
as the denominator goes to 0 (limr→ Rπ sin Rr = sin Rπ

R = sin(π ) = 0). The angular width is minimized
when the object lies on the equator, where r = Rπ dl
2 and dθ = R , and this angular width increases to infinity
as the object approaches the antipode of the sphere. For an object at the antipode, all lines of sight from
the observer lead to the object, and it therefore fills the horizon.

8
3.3 The Earth isn’t flat!
We choose our coordinate system such that the origin is located in the center of the sphere, and we as the
oberver are located at the north pole. The circle drawn on the sphere is at a fixed distance r from us on
the north pole, and therefore each point on the sphere is at a fixed polar angle θ. The radius of this circle
as measured as an outside observer (not located on the sphere) is R sin θ. The circumference of the circle

R sin θ θ
R

Figure 1: Illustration of setup of Problem 3.3

is therefore given by:


C = 2πR sin θ (3.8)
Now, we recall that the arclength r is related to the polar angle and sphere radius by:

r = Rθ. (3.9)

So substituing this into the circumference, we obtain:


 
r
C = 2πR sin (3.10)
R

which was the claimed formula. For the second part of the question, we consider that for a flat space, the
circumference would be measured to be:
C f = 2πr. (3.11)
In the limit r  R, we can Taylor expand the circumference as given in Eq. (3.10) to obtain that:
!
r r3 πr3
Cs ≈ 2πR − 3 = 2πr − 2 (3.12)
R 6R 3R

Hence the difference between the circumference measured in Euclidean space versus a sphere would be
given by:
πr3 πr3
C f − Cs ≈ 2πr − (2πr − 2
) = . (3.13)
3R 3R2
We can measure distances within an error of ±1m, so to convince ourselves that Earth is spherical rather
than flat, we require that the circumference differnece be:

C f − Cs > 1m (3.14)

So approximately:
πr3
> 1m (3.15)
3R2

9
Rearranging, we have: r
3 3mR2
r> (3.16)
π
And substituing R = 6400km we get:
r > 34km (3.17)

3.4 Area Bounds for Equilateral Triangles


Case 1: κ = +1. No, we cannot draw an equilateral triangle of arbitrarily large surface area A in this case.
A simple counterargument is that a sphere has surface area 4πR2 , so immediately no triangle with area
larger than that is possible. For a calculation of the maximum area of a triangle, we require a thoughtful
argument. First, recall the equation for the sum of three angles of a triangle of area A on a surface of a
sphere of radius R:
A
α+β+γ = π+ 2 (3.18)
R
For an equiliateral triangle, α = β = γ, so:

A
3α = π + (3.19)
R2
WLOG, we can choose our coordinate system that the first point of the triangle is on the north pole, the
second point is a distance r from the north pole on the sphere with azimuthal angle φ = 0, and the third
point is also a distance r from the north pole on the sphere with azimuthal angle φ = α.

α
α

Figure 2: Illustration of the triangle drawn for the spherical case of Problem 3.4.

We now recall the area element on the sphere is given as:

dA = Rdθ · R sin θdφ = R2 dθdφ (3.20)

Where Rdθ is the polar distance and R sin θdφ is the azimuthal distance. With the setup as described, the
area of the triangle may be calculated via integration. We integrate from 0 ≤ θ ≤ Rr (recalling that r = Rθ)
and from 0 ≤ φ ≤ α:
ZZ Z r/R Z α
2
A= dA = R sin θdθ dφ (3.21)
triangle 0 0

Carrying out these two easy integrals, we obtain:


 !
2 r
A=R 1 − cos α (3.22)
R

10
Isolating for α, we obtain:
A
α=   (3.23)
r
R2 1 − cos R

Plugging this into (3.19), we get:


A A
3 
r
  = π + R2 (3.24)
R2 1 − cos R

Isolating for A we have:


1 − cos Rr

A= πR2 (3.25)
2 + cos Rr


The largest that r can possibly be is when r = πR (the base of the triangle lies on the south pole), and so:
 
1 − cos πR
R 1 − (−1)
Amax = πR2   = πR2 = 2πR2 (3.26)
2 + cos πR 2−1
R

In other words, the largest an equilateral triangle can be in this case is being half of the sphere!
Case 2: κ = 0. We can draw an equilateral triangle of arbitrarily large surface area A in this case. On a
flat plane, there is no upper bound to how large you want to draw your shapes!
Case 3: κ = −1. No, we cannot draw an equilateral triangle of arbitrarily large surface area A in this case.
Consider the equation that gives the sum of three angles of a triangle on a 2-D surface of uniform negative
curvature (with radius of curvature R):

A
α+β+γ = π− (3.27)
R2
For an equilateral triangle, α = β = γ so:
A
3α = π − (3.28)
R2
α cannot be negative if we are to have a physical shape, so:

A
0≤π− (3.29)
R2
Rearranging, we obtain an upper bound on the area of an equilateral triangle for a 2-D surface of uniform
negative curvature:
A ≤ R2 π (3.30)
Which gives us a maximum:
Amax = R2 π (3.31)
in this limit, note that the angles of the triangle α approach zero.

3.5 Equivalent Metrics


We wish to show the equivalence of the metrics:

dl 2 = dx2 + dy2 + dz2 (3.32)

and: h i
dl 2 = dr2 + r2 dθ 2 + sin2 θdφ2 . (3.33)

11
Making the substitutions x = r sin θ cos φ, y = r sin θ sin φ, and z = r cos θ into (3.32), we have:

dl 2 = (d(r sin θ cos φ))2 + (d(r sin θ sin φ))2 + (d(r cos θ ))2 (3.34)

Liberally applying the product and chain rules, we obtain:

dl 2 = (sin θ cos φdr + r cos θ cos φdθ − r sin θ sin φdφ)2


+ (sin θ sin φdr + r cos θ sin φdθ + r sin θ cos φdφ)2
+ (cos θdr − r sin θdθ )2 (3.35)

Now we can expand this expression:

dl 2 = sin2 θ cos2 φdr2 + r2 cos2 θ cos2 φdθ 2 + r2 sin2 θ sin2 φdφ2


+ 2r sin θ cos θ cos2 φdrdθ − 2r sin2 θ sin φ cos φdrdφ − 2r2 sin θ cos θ sin φ cos φdθdφ
+ sin2 θ sin2 φdr2 + r2 cos2 θ sin2 φdθ 2 + r2 sin2 θ cos2 φdφ2
+ 2r sin θ cos θ sin2 φdrdθ + 2r sin2 θ sin φ cos φdrdφ + 2r2 sin θ cos θ sin φ cos φdθdφ
+ cos2 θdr2 − 2r sin θ cos θdrdθ + r2 sin2 θdθ 2 (3.36)

Now grouping like terms, we have:


h i h i
dl 2 = dr2 sin2 θ cos2 φ + sin2 θ sin2 φ + cos2 θ + r2 dθ 2 cos2 θ cos2 φ + cos2 θ sin2 φ + sin2 θ
h i h i
+ r2 dφ2 sin2 θ sin2 φ + sin2 θ cos2 φ + 2rdrdθ sin θ cos θ cos2 φ + sin θ cos θ sin2 φ − sin θ cos θ
h i
+ 2rdrdφ − sin2 θ sin φ cos φ + sin2 θ sin φ cos φ + 2r2 dθdφ − sin θ cos θ sin φ cos φ + sin θ cos θ sin φ cos φ
 

(3.37)

We can see the last line of terms all equals to zero (the brackets evaluate to zero, so):
h i h i
dl 2 = dr2 sin2 θ cos2 φ + sin2 θ sin2 φ + cos2 θ + r2 dθ 2 cos2 θ cos2 φ + cos2 θ sin2 φ + sin2 θ
h i h i
+ r2 dφ2 sin2 θ sin2 φ + sin2 θ cos2 φ + 2rdrdθ sin θ cos θ cos2 φ + sin θ cos θ sin2 φ − sin θ cos θ (3.38)

Now redrawing some brackets:


h i h i
dl 2 = dr2 sin2 θ (cos2 φ + sin2 φ) + cos2 θ + r2 dθ 2 cos2 θ (cos2 φ + sin2 φ) + sin2 θ
h i h i
+ r2 dφ2 sin2 θ (sin2 φ + cos2 φ) + 2rdrdθ sin θ cos θ (cos2 φ + sin2 φ) − sin θ cos θ (3.39)

Now applying the cos2 φ + sin2 φ = 1 identity to each term:


h i h i h i
dl 2 = dr2 sin2 θ + cos2 θ + r2 dθ 2 cos2 θ + sin2 θ + r2 dφ2 sin2 θ + 2rdrdθ [sin θ cos θ − sin θ cos θ ]
(3.40)
The last term vanishes, and to the first two terms we can apply the identity cos2 θ + sin2 θ = 1:

dl 2 = dr2 + r2 dθ 2 + r2 sin2 θdφ2 (3.41)

And redrawing some brackets: h i


dl 2 = dr2 + r2 dθ 2 + sin2 θdφ2 (3.42)
We identify this with (3.33) and hence we conclude.

12
4 Cosmic Dynamics
4.1 Does the Cosmological Constant Affect Planetary Motion?
In a sphere of radius 1AU, we have:

4 4
EΛ = eλ V = eλ πR3 = 5200MeV m−3 π (1.5 × 1011 m)3 = 1.18 × 1025 J (4.1)
3 3

Now calculating the rest energy of the sun, we have:

E = M c2 = 1.99 × 1030 kg(3.0 × 108 ms−1 )2 = 1.79 × 1047 J (4.2)

We see a difference of 23 orders of magnitude; we conclude that the cosmological constant does not have
a significant effect on the motion of planets within the Solar system.

4.2 Perturbing Einstein’s Static Universe


If Λ = 4πGρ, then the acceleration equation says:

ä 4πG Λ 4πGρ 4πGρ


= − 2 (e + 3P) + = − + =0 (4.3)
a 3c 3 3 3
Where in the second equality we use that this hypothetical universe is filled solely with matter, so P ≈ 0.
Now, if some of this matter gets converted to radiation, we have that Pr = 13 e > 0, so Ptot > 0 and hence:

ä 4πG Λ 4πGPtot
= − 2 (e + 3P) + = − <0 (4.4)
a 3c 3 3
Note that while the pressure increases, the energy density e remains unchanged (the energy just converted
form) so the energy density term and the cosmological constant term cancel out like before. From this, we
conclude that ä < 0, and so the universe contracts . The extra gravitational pressure from the radiation
causes the universe to collapse; this shows that Einstein’s static universe model isn’t great, as even his
universe with just one star would trigger a runaway collapse.

4.3 How Large is Einstein’s Static Universe?


From Eq. 4.73 in Ryden, we have that in Einstein’s static universe has radius of curvature:

c
R0 = = 2 × 1026m ∼ 7Gpc (4.5)
2(πGρ)1/2

If a photon were to circumnavigate this universe, it would take time:

2πR0
T= = 4 × 1018 s ∼ 132Gyr (4.6)
c

Which is longer than the age of the universe!

13
4.4 Baseballs and Critical Density
The current critical density is given by Ryden Eq. 4.32 to be:

3
ρc,0 = H 2 = 8.7 × 10−27 kg m−3 (4.7)
8πG 0
We set the density of baseballs to be equal to the critical density:

ρc,0 = ρbb = mbb nbb (4.8)

Where nbb is the number density of the baseballs. Rearranging, we get:

ρc,0 8.7 × 10−27 kg m−3


nbb = = = 6.0 × 10−26 m−3 (4.9)
mbb 0.145kg

Given this density of baseballs, we can use Ryden Eq. 2.2 to solve for the average distance we could see
before having our line of sight intersected by a baseball:

1 1
λ= 2
= −
= 3.90 × 1027 m ≈ 126000Mpc (4.10)
nbb πrbb (6.0 × 10 m 3 )π (0.0369m)2
− 26

The fact that we can see galaxies at a distance ∼ c/H0 ∼ 4000Mpc does not give us a useful upper
bound on the density of intergalatic baseballs in this case (we see that the line of sight from the current
calculation assuming critical density of baseballs is ∼2 orders of magnitude larger than what we can
actually see already). However, for completeness we calculate what upper bound this does give on the
density of intergalatic baseballs:

1 1
nbb < 2
= = 1.93 × 10−24 m−3 (4.11)
λπrbb (4000Mpc)(π (0.0369m)2 )

4.5 Equation of State for Gases


The energy per-particle is given by:
E = (mc2 + h2 c2 /λ2 )1/2 (4.12)
And the total energy density of a gas of particles is given by:

e = nE (4.13)

Combining the two, we have:


e = n(mc2 + h2 c2 /λ2 )1/2 (4.14)
Since n is the number density, we can write it as:

N N
n= = (4.15)
V k 1 a3
Where N is the number of particles in the gas (we assume this does not change, i.e. that no particles
are created or destroyed), and V = k1 a3 is the volume of the expanding universe (proportional to a3 ).
Furthermore, we can write λ = k2 a as the wavelength is linear in the scale factor. Putting this into (4.14)
we have:
N 2 h2 c2 1/2
e= ( mc + ) (4.16)
k 1 a3 k22 a2

14
Now we recall the fluid equation:

ė + 3 (e + P) = 0. (4.17)
a
Substituting the equation of state:
P = we (4.18)
into the fluid equation, we have:

ė + 3 e(1 + w) = 0 (4.19)
a
Solving for w, we have:
−ė
w= −1 (4.20)
3 ȧa e
We will have to take the time derivative of (4.16) to substitute into (4.20). Noting that the only time-
dependent parameter in e is a, we take the derivative (using the quotient rule and chain rule):
 
N h2 c2 2 2
2 h2 c2 1/2
− 2 k 2 a3
ȧk1 a3 − 3Nk1 a2 ȧ(mc2 + kh2 ac2 )1/2
(mc + ) 2 2
k22 a2
ė = (4.21)
k21 a6
Simplifying slightly:
− Nc2 ȧ(3k22 ma2 + 4h2 )
ė = (4.22)
h2 c2 1/2
k1 k22 a6 (mc2 + k22 a2
)
Substituting (4.16) and (4.22) into (4.20) we have:
Nc2 ȧ(3k22 ma2 +4h2 )
2 2
k1 k22 a6 (mc2 + h2 c2 )1/2
k2 a
w=
h2 c2 1/2
−1 (4.23)
3 ȧa k Na3 (mc2 + k22 a2
)
1

Cancelling terms in the numerator and denominator, we have:


c2 (3k22 ma2 + 4h2 )
w=
h2 c2
−1 (4.24)
3k22 a2 (mc2 + k22 a2
)

Expanding out terms in the numerator and denominator:


3mc2 k22 a2 + 4h2 c2
w= −1 (4.25)
3mc2 k22 a2 + 3h2 c2
In the highly relativistic limit, we have a → 0 and p → 0. Note that while p does not appear explicitly
in the above equation, a → 0 implies p → ∞ under the linear relationship of the scaling factor with λ, as
k2 a = λ = h/p. In any case, taking a → 0 in the above expression, we have:

3mc2 k22 a2 + 4h2 c2 4h2 c2 4 1


wrel = lim a→0 w = lim − 1 = −1 = −1 = (4.26)
a→w 3mc2 k2 a2 + 3h2 c2 2
3h c 2 3 3
2

This was precisely the claimed value. Now in the highly non-relativistic limit, we have a → ∞ and p → 0.
We again take the limit of a → ∞ in (4.25) to obtain:

3mc2 k22 a2 + 4h2 c2 3mc2 k22


wnonrel = lim w = lim − 1 = −1 = 1−1 = 0 (4.27)
a→∞ a→∞ 3mc2 k2 a2 + 3h2 c2 3mc2 k22
2

which is again the desired value.

15
A Force-Based Derivation of the Newtonian Friedmann Equation

Now let’s derive the “acceleration equation” (sometimes called “Friedmann’s other equation”!) for
the whole Universe from simple Newtonian physica. Imagine a sphere of constant density ρ(t) and
radius r, with a test mass m at its edge. Write down the equation of motion for the test mass under
the gravitational pull of hte sphere. Now use the idea that the physical radius can be written as
comoving radius times scale factor, i.e. r ≡ a(t) x. you should find that you can derive an equation
for a which doesn’t depend on x or on m! In other words, the sphere that oyu used inthe first place
has dissapeared and your equation of motion has ended up being for the scale factor itself. [Note
that you’re not being asked to solve this equation, just to derive it!]

The mass of the sphere is given by:

4
M (t) = ρ(t)V (t) = ρ(t) πr (t)3 (4.28)
3
The distance from the center of the sphere to the test mass is just r (t) (the test mass is on the surface), so
using Newton’s second law and Newton’s law of universal gravitation, we have:

− GM(t)m
mr̈ (t) = F = (4.29)
r ( t )2

Substituting M (t) from (4.28), we have:

− Gρ(t) 34 πr (t)3 m 4
mr̈ (t) = 2
= − Gρ(t) πa(t)m (4.30)
r (t) 3

Cancelling out m from both sides and replacing r (t) with a(t) x we have:

4
x ä(t) = − Gρ(t) πa(t) x (4.31)
3
The xs cancel on both sides, and dividing both sides by a(t) we get:

ä(t) 4
= − Gρ(t) π (4.32)
a(t) 3

16
5 Model Universes
5.1 Redshift in single-component universes
We can take Eq. 5.47 in Ryden as our starting point:
 2/(3+3w)
a ( t0 ) t0
1+z = = (5.1)
a(te ) te
Taking the derivative w.r.t. t0 of both sides of this equation, we obtain:
dt0 dte
ȧ(t0 ) a(te ) − ȧ(te ) a ( t0 )
dz dt0 dt0
= (5.2)
dt0 a ( t e )2
There’s a variety of terms to process here. First, combining Ryden Eqs. 5.39 and 5.42 we get:
 2/(3+3w)
t
a(t) = (5.3)
t0
Taking the time derivative, we have:
 2/(3+3w)
2 t 1 2 a(t)
ȧ(t) = = (5.4)
3 + 3w t0 t 3 + 3w t
We also have that:
dt0
=1 (5.5)
dt0
dte
The final quantity to determine is dt0 . To solve for this, we recall Ryden 3.59:
Z te +λe /c Z t0 +λ0 /c
1 1
dt = (5.6)
a(te ) te a ( t0 ) t0

Since λ/c  1, we can approximate λe /c ∼ dte , λ0 /c ∼ dt0 to get:


Z te +dte Z t0 +dt0
1 1
dt = dt (5.7)
a(te ) te a ( t0 ) t0

Which solving the integrals tells us that:


dte dt0
= (5.8)
a(te ) a ( t0 )
Which we can rearrange to obtain:
dte a(te ) 1
= = (5.9)
dt0 a ( t0 ) 1+z
Combining the three obtained relations of (5.4), (5.5), (5.9) and substituting this into (5.2) we have that:
2 a ( t0 ) 2 a(te ) 1
dz 3+3w t0 a ( te ) − 3+3w te 1+z a ( t0 )
= (5.10)
dt0 a ( t e )2
a ( t0 )
Factoring and using that a(te )
= 1 + z from (5.1) this becomes:
 
dz 2 1 1
= (1 + z ) − (5.11)
dt0 3 + 3w t0 te

17
Substituting te = t0 (1 + z)−(3+3w)/2 from (5.1) we have:
 
dz 2 1 (3+3w)/2 1
= (1 + z ) − (1 + z ) (5.12)
dt0 3 + 3w t0 t0

Now using Ryden Eq. 5.42:


2
t0 = H −1 (5.13)
3 + 3w 0
We obtain:
dz
= H0 (1 + z) − H0 (1 + z)(3+3w)/2 (5.14)
dt0
dz
which was the desired relation. The observed redshift increases in time if dt0 > 0, so rearranging the
above expression we have:

H0 (1 + z) − H0 (1 + z)(3+3w)/2 > 0 =⇒ (1 + z) > (1 + z)(3+3w)/2 (5.15)

And since 1 + z ≥ 1, the LHS is greater than the RHS if (3 + 3w)/2 < 1, i.e. if:

1
w<− . (5.16)
3

5.2 Redshift Change Timescale for Flat Matter-Only Universe


In a flat matter-only universe, we can use the result from 5.1 with w = 0 to obtain that:

dz
= H0 (1 + z) − H0 (1 + z)3/2 (5.17)
dt0

We want to find the time dt0 for the galaxy to change by dz −6 −6


z = −10 , and since z = 1 to start dz = −10 .
Rearranging the above expression to solve for dt0 we have:

dz 1
dt0 = (5.18)
H0 (1 + z) − (1 + z)3/2

So plugging in dz = −10−6 , z = 1, and H0 = 68kms−1 Mpc−1 we numerically get:

dt0 ≈ 5.49 × 1011 s ≈ 17400yrs (5.19)

5.3 Present age of universe for positively curved matter-only universe


We recall the parametric solutions for the scale factor a and the time t for a positively curved universe
filled only with matter:
1 Ω0
a(θ ) = (1 − cos θ ) (5.20)
2 Ω0 − 1
1 Ω0
t(θ ) = (θ − sin θ ). (5.21)
2H0 (Ω0 − 1)3/2
In the present day, a = 1, so solving (5.20) for θ we have:
!
2( Ω0 − 1) 2 − Ω0
 
−1
θ = cos 1− = cos−1 (5.22)
Ω0 Ω0

18
Substituting this into (5.21) we have:
!
Ω0 2 − Ω0 2 − Ω0
  
1
t0 = (cos−1 − sin cos −1
) (5.23)
2H0 (Ω0 − 1)3/2 Ω0 Ω0

The last term looks complicated, but we can view cos−1 2−ΩΩ0 0 as the angle for a triangle with hypotenuse
 

Ω0 and adjacent side 2 − Ω0 . The sine of this will therefore be the ratio of the opposite side and the
hypotenuse of this triangle. The length of the opposite side is given (by Pythagoras) as:
q p
Ω20 − (2 − Ω0 )2 = 2 Ω0 − 1 (5.24)
 
2− Ω0

So sin cos−1 Ω0 is given as:

! √
2 − Ω0 2 Ω0 − 1

−1
sin cos = . (5.25)
Ω0 Ω0

So substituting this into (5.23) we get:



Ω0 2 − Ω0 2 Ω0 − 1
 
1
H0 t0 = (cos−1 − ) (5.26)
2 (Ω0 − 1)3/2 Ω0 Ω0

Which distributing the product we get:

Ω0 2 − Ω0
 
1 1
H0 t0 = cos−1 − (5.27)
2 (Ω0 − 1)3/2 Ω0 Ω0 − 1

which was the desired expression. A plot of t0 vs. Ω0 for 1 ≤ Ω0 ≤ 3 is given below.

t0 vs Omega0
0

20
t0 (s * MPc/km)

40

60

80

100
1.00 1.25 1.50 1.75 2.00 2.25 2.50 2.75 3.00
Omega0

Figure 3: Plot of t0 vs. Ω0 for 1 ≤ Ω0 ≤ 3.

19
5.4 Present age of universe for negatively curved matter-only universe
We recall the parametric solutions for the scale factor a and the time t for a negatively curved universe
filled only with matter:
1 Ω0
a(η ) = (cosh η − 1) (5.28)
2 1 − Ω0
1 Ω0
t(η ) = (sinh η − η ). (5.29)
2H0 (1 − Ω0 )3/2
In the present day, a = 1, so solving (5.28) for η we have:
!
−1 2 (1 − Ω 0 ) −1 2 − Ω 0
 
η = cosh + 1 = cosh (5.30)
Ω0 Ω0

Substituting this into (5.29) we get:


!
Ω0 2 − Ω0 2 − Ω0
  
1
t0 = (sinh cosh−1 − cosh−1 ) (5.31)
2H0 (1 − Ω0 )3/2 Ω0 Ω0
  √
Now, using that sinh cosh−1 x = x2 − 1, the first term of the above expression becomes:

! !1/2 !1/2
2 − Ω0 (2 − Ω0 )2 4 − 4Ω0

2
sinh cosh−1 = −1 = = (1 − Ω0 )1/2 (5.32)
Ω0 Ω20 Ω20 Ω0

So substituting this result we obtain:


!
Ω0 2 − Ω0

1 2
H0 t0 = (1 − Ω0 )1/2 − cosh−1 (5.33)
2 (1 − Ω0 )3/2 Ω0 Ω0

Distributing the terms, we obtain the desired result:

Ω0 2 − Ω0
 
1
H0 t0 = − cosh−1 . (5.34)
1 − Ω0 2(1 − Ω0 )3/2 Ω0

Finally, a plot of t0 vs. Ω0 for 0 ≤ Ω0 ≤ 1 is given below.

5.5 Phantom Energy and the Big Rip


The supposed “phantom energy” with equation of state parameter w p < −1 would have energy density:

e p ( a) = e p,0 a−3(1+w p ) . (5.35)

Comparitively, matter has energy density:

em ( a) = em,0 a−3 . (5.36)

At equality, we have that:


e p ( a) e p,0 a−3(1+w p ) Ω p,0 1
1= = = (5.37)
em ( a ) em,0 a − 3 Ωm,0 a3w p
Since Ω p,0 = 1 − Ωm,0 , we have:
1 − Ωm,0 1
1= (5.38)
Ωm,0 a3w p

20
t0 vs Omega0
0.014

0.013
t0 (s * MPc/km)
0.012

0.011

0.010
0.0 0.2 0.4 0.6 0.8 1.0
Omega0

Figure 4: Plot of t0 vs. Ω0 for 0 ≤ Ω0 ≤ 1.

Which we can rearrange to solve for the scale factor amp where equality holds:
!1/3w p
1
amp = −1 (5.39)
Ωm,0

The Friedmann equation in this universe reads:


H2 Ω Ω p,0 Ω 1−Ω
2
= m,0
3
+ 3(1+w ) = m,0
3
+ 3(1+wm,0) (5.40)
H0 a a p a a p

In the limit a  amp , the first term becomes negligible and hence:

H2 1−Ω
≈ 3(1+wm,0) (5.41)
H02 a p

We now rearrange this equation to set up for the integration:



(1 − Ωm,0 )1/2
a
≈ =⇒ H0 dt ≈ a3(1+w p )/2−1 (1 − Ωm,0 )−1/2 da (5.42)
H0 a3(1+w p )/2
Integrating from t0 to trip , corresponding to from a(t0 ) = 1 to a(trip ) = ∞ on the RHS, we have:
Z trip Z ∞
H0 dt ≈ a3(1+w p )/2−1 (1 − Ωm,0 )−1/2 da (5.43)
t0 1

Carrying out the integral, we get:



2
H0 (trip − t0 ) ≈ a3(1+w p )/2 (1 − Ωm,0 )−1/2 (5.44)
3(1 + w p )
1

Since w p < −1, the RHS goes to 0 at a = ∞, so:


2
H0 (trip − t0 ) ≈ − (1 − Ωm,0 )−1/2 (5.45)
3(1 + w p )

21
Which we can write as:
2
H0 (trip − t0 ) ≈ (1 − Ωm,0 )−1/2 (5.46)
3 1 + wp

which was the desired result. Finally, with H0 = 68km s−1 Mpc−1 , Ωm,0 = 0.3, and w p = −1.1, we can
solve numerically for the time remaining until the “Big Rip” to be:

trip − t0 ≈ 115.5Gyr (5.47)

5.6 Pulling an Einstein


In this universe with matter and dark energy, the acceleration equation reads:
ä 4πG 4πG
= − 2 (e + 3P) = − 2 (em + eq + 3wq eq ) (5.48)
a 3c 3c
Where we have used that P = (0)em = 0 for matter and P = wq eq for the dark energy. To have a static
universe with ä = 0, it must follow that:

em = −(1 + 3wq )eq (5.49)

The Friedmann equation reads:


8πG κc2
H ( t )2 = 2
( em + eq ) − 2 (5.50)
3c R0 a ( t )2
Since we have a static universe, ȧ = 0 and hence H (t) = 0. Furthermore, using that em = −(1 + 3wq )eq
from earlier, the Friedmann equation becomes:

8πG κc2
0= 2
(−3wq )eq − 2 (5.51)
3c R0 a ( t )2

Since −1 < wq < −1/3, wq is negative, and hence the only way the above equality is satisfied is if κ = 1 so
the positive term can be cancelled by the second term. So in this scenario, the universe is positively curved .
We rearrange the above equation to solve for the radius of curvature R0 when a(t) = 1:
v
c4
u
R0 = t (5.52)
u
8πG wq eq

5.7 Big Crunch


For a positively curved matter only universe, we recall the parametric solutions for a(θ ), t(θ ) (Ryden Eqs.
5.90/5.91):
1 Ω0
a(θ ) = (1 − cos θ ) (5.53)
2 Ω0 − 1
1 Ω0
t(θ ) = (θ − sin θ ) (5.54)
2H0 (Ω0 − 1)3/2
The big bang occurs at θ = 0, and the Big crunch at θ = 2π. Given the parametric solutions above, he time
between these can be computed as (Ryden Eq. 5.92):
π Ω0
tcrunch = (5.55)
H0 (Ω0 − 1)3/2

22
Furthermore, we can solve for the present t0 at which a(θ0 ) = 1:
!
1 Ω0 2( Ω0 − 1)
 
2
a ( θ0 ) = 1 = (1 − cos θ0 ) =⇒ θ0 = arccos 1 − = arccos −1 (5.56)
2 Ω0 − 1 Ω0 Ω0

Therefore the time between Dr. Niwde’s obsdrvations at t = t0 = t(θ0 ) and the final big crunch is given
by:
π Ω0 1 Ω0
∆t = tcrunch − t(θ0 ) = − (θ0 − sin θ0 ) (5.57)
H0 (Ω0 − 1) 3/2 2H0 (Ω0 − 1)3/2
Or more concisely:

 
1 1
∆t = π − (θ0 − sin θ0 ) (5.58)
H0 (Ω0 − 1)3/2 2

Where θ0 is given in (5.56). To determine the highest amplitude blueshift, we recall the redshift-scale factor
relation 1 + z = 1a , so:
1
z = −1 (5.59)
a
If we want to minimize z (i.e. have it be the most negative/highest magnitude blueshift), we want to
maximize a. We can determine this from the Friedmann equation (a la Ryden Eq. 5.86) but it also can
easily be read off from the parametric solution above to be:

Ω0
amax = (5.60)
Ω0 − 1

which is attained at θ = π. So, the highest amplitude blueshift that Dr. Niwde can observe is at:

Ω0 − 1 1
zblue, max = −1 = − (5.61)
Ω Ω0

At this blueshift, as stated previously we have θ = π. So this occurs at:

1 Ω0 π Ω0
tblue, max = t(π ) = (π − sin π ) = (5.62)
2H0 (Ω0 − 1) 3/2 2H0 (Ω0 − 1)3/2

So the lookback time is:


1 Ω0 π Ω0
t0 − tblue, max = (θ0 − sin θ0 ) − (5.63)
2H0 (Ω0 − 1)3/2 2H0 (Ω0 − 1)3/2

Or with some factoring:

1 Ω0
t0 − tblue, max = (θ0 − sin θ0 − 1) (5.64)
2H0 (Ω0 − 1)3/2

5.8 Big Bounce


The Friedmann equation in such a universe would read:

H2 1 − Ω0
2
= Ω0 + (5.65)
H0 a2

23
At the point where a has an extrema, H (t) = 0 and so:
1/2
1 − Ω0 Ω0 − 1

0 = Ω0 + =⇒ abounce = (5.66)
a2 Ω0

Rearranging the Friedmann equation, we have:


s
ȧ p a2 da H0
q
= H0 Ω0 1 − bounce =⇒ = √ a2 − a2bounce (5.67)
a a2 dt Ω0
Integrating, we obtain:
Z t0 Z a
1
da0
p
Ω0 H0 dt = q (5.68)
tbounce abounce a 02 − a2bounce
Integrating both sides (making use of an integral table for the RHS), we have:
 
p a
Ω0 H0 (t − tbounce ) = arccosh (5.69)
abounce
Which we can rearrange to obtain:
p 
a(t) = abounce cosh Ω0 H0 (t − tbounce ) (5.70)

At t0 we have that a(t0 ) = 1 by convention, so solving for t0 − tbounce using (5.69) we have:
1/2 !
Ω0

1
t0 − tbounce = √ arccosh (5.71)
H0 Ω0 Ω0 − 1

5.9 Ωm,0 for t = H0−1


In a spatially flat and matter + cosmological constant filled universe, the Friedmann equation can be
integrated to yield the analytic solution relating t and a (Ryden Eq. 5.101):
 s 
 3/2  3
2 a a
H0 t = p ln  + 1+  (5.72)
3 1 − Ωm,0 amΛ amΛ

Where amΛ is defined by:


!1/3
Ωm,0
amΛ = (5.73)
1 − Ωm,0
At the present time, we have t = t0 , and a = a(t0 ) = 1 by convention. Substituting this, as well as (5.73)
into (5.72) we get: s 
s
2 1 − Ωm,0 1 − Ωm,0 
H0 t0 = p ln  + 1+ (5.74)
3 1 − Ωm,0 Ωm,0 Ωm,0

In order to have t0 = H0−1 we require the RHS of the above equation to exactly equal one:
s s 
2 1 − Ωm,0 1 − Ωm,0 
1= p ln  + 1+ (5.75)
3 1 − Ωm,0 Ωm,0 Ωm,0

24
Rewriting the last term:
s s  " #
Ω 1 + 1 − Ωm,0
p
2 1 − m,0 1 2
1= p ln  + = p ln (5.76)
3 1 − Ωm,0 Ωm,0 Ωm,0 3 1 − Ωm,0 Ωm,0
p

 √ 
1+ 1− x 2
Using the hyperbolic secant identity of arcsech x = ln x , we have:

2
arcsech( Ωm,0 )
p
1= p (5.77)
3 1 − Ωm,0
Or:
1 − Ωm,0
p
3
= arcsech( Ωm,0 )
p
(5.78)
2
This equation can now be solved numerically for Ωm,0 to find:

Ωm,0 ≈ 0.263 (5.79)

5.10 Amounts in the Benchmark model


In the Benchmark model, we have (from Ryden Table 5.2) that:

Ωm,0 = 0.31, Ωγ,0 = 5.35 × 10−5 , Ωbary,0 = 0.048 (5.80)


e(t)
We that Ω = e (t) where eC is the critical density, which is currently ec,0 7.8 × 10−10 J m−3 (up to uncer-
c
tainty). Therefore obtaining the energy density for photons in the benchmark model, we have:

eγ,0 = ec,0 Ωγ,0 = 4.17 × 10−14 J m−3 (5.81)


It will be more convenient to solve for the mass density for the matter and baryons:

ρm,0 = ec,0 Ωγ,0 /c2 = 2.69 × 10−27 kg m−3 , ρbary,0 = ec,0 Ωbary,0 /c2 = 4.16 × 10−28 J m−3 (5.82)
In addition, in the Benchmark model we have a finite horizon distance (Ryden Eq. 5.115):

dhor (t0 ) = 14000Mpc = 4.33 × 1026 m. (5.83)


Therefore the volume of the universe within the horizon distance is:
4
Vhor (t0 ) = πd (t0 )3 = 3.39 × 1080 m3 (5.84)
3 hor
Now, solving for the total mass of all matter within our horizon, we have:

Mm = Vhor (t0 )ρm,0 = 9.12 × 1053 kg (5.85)

Next, solving for the total amount of energy of photons within the horizon distance, we have:

Ephotons = Vhor (t0 )eγ,0 = 1.41 × 1067 J (5.86)

Finally, we solve for the number of baryons in the universe. This will be the total mass of baryons within
the universe divided by the mean mass per Baryon, which we take to be the mass of a proton m p . Therefore
the total number of baryons within the horizon can be solved for as:

Nbary = Vhor (t0 )ρbary,0 /m p = 8.45 × 1079 baryons (5.87)

25
Time and Scale Factor for Matter-Only Universe

Let’s make sure we can work through the mathematical steps for a closed matter-only (i.e. ‘Matter
+ Curvature’) universe. Try to do this without looking up the book for every step! Start by writing
the Friedmann equation for this case, using a rather than 1 + z, and with the single parameter Ω0
(where this implicitly means ‘matter’, here). Thus write an integral expression for t. It may not look
trivial to solve this for t( a) or a(t), but you should be able to show that the following parametric
solution works [a ‘parametric solution’ means that you can write doesn y(φ) and x (φ), both in
terms of a parameter φ, even if an explicit expression for y( x ) is hard or impossible]:

1 Ω0
a(θ ) = (1 − cos θ ) (5.88)
2 Ω0 − 1
1 Ω0
t(θ ) = (θ − sin θ ). (5.89)
2H0 (Ω0 − 1)3/2
To be clear: you are being asked to show that in the appropriate Friedmann equation, the LHS
is equal to the RHS if you assume the above solution (or you could solve the integral, but that’s
harder!). [You might have to do a bit of rearranging of trig functions - but perservere, because it
works!] Lastly, by letting θ run from 0 to 2π, sketch a vs. t for this model.

The Friedmann equation in this case reads:

H2 Ω 1 − Ω0
2
= 30 + (5.90)
H0 a a2

Now, since H = ȧa , we can rearrange this to obtain that:

ȧ2 Ω
2
= ( 0 + 1 − Ω0 ) (5.91)
H0 a

Or in other words: r
da Ω0
= H0 + 1 − Ω0 (5.92)
dt a
So solving for t we integrate:
da0
Z t Z a
0
H0 t = dt = q . (5.93)
0 0 Ω0
a + 1 − Ω0

Now, we verify that (5.91) holds for the parametric solution given by (5.88) and (5.89). First, we determine
what ȧ is given this solution. By the chain rule, we have that:
da da dθ da
ȧ = = = θ̇ (5.94)
dt dθ dt dθ
da
Solving for dθ by differentiating (5.88) we have:
da 1 Ω0
= sin θ. (5.95)
dθ 2 Ω0 − 1
And solving for θ̇ by implicitly differentiating (5.89) we have:

1 Ω0 2H0 (Ω0 − 1)3/2


1= ( θ̇ − θ̇ cos θ ) =⇒ θ̇ = (5.96)
2H0 (Ω0 − 1)3/2 Ω0 (1 − cos θ )

26
Hence, we find that:
H0 (Ω0 − 1)1/2
ȧ = sin θ (5.97)
1 − cos θ
Evaluating the LHS of (5.91), we have:

ȧ2 Ω0 − 1
= sin2 θ (5.98)
H02 (1 − cos θ )2

Evaluating the RHS of (5.91), we have:

Ω0 Ω0
+ 1 − Ω0 = ( 1 Ω0
+ 1 − Ω0 ) (5.99)
a
2 Ω1 −1 (1 − cos θ )

 2
1−cos θ 1−cos θ
Multiplying the first term by 1 = 1−cos θ and the second term by 1 = 1−cos θ we obtain:

Ω0 2( Ω0 − 1) Ω0 − 1
+ 1 − Ω0 = (1 − cos θ ) − (1 − cos θ )2 (5.100)
a (1 − cos θ )2 (1 − cos θ )2

Which after some expanding and cancellation, becomes:

Ω0 Ω0 − 1
+ 1 − Ω0 = (1 − cos2 θ ) (5.101)
a (1 − cos θ )2

Using the famous trig identity sin2 θ + cos2 θ = 1, we can identify (5.98) with (5.101) to conclude that this
is indeed the correct solution.
For the plot, we observe that the given parametric equations are exactly of those for a cycloid, so the
curve of t(θ ) vs. a(θ ) will be exactly that it is displayed below.

t( ) vs. a( )
2.00
1.75
1.50
1.25
1.00
t( )

0.75
0.50
0.25
0.00
0 1 2 3 4 5 6
a( )

Figure 5: Plot of t(θ ) vs. a(θ ) for θ ∈ [0, 2π ). Ω0 = 2 and H0 = 1 were chosen for convenience of plotting.

27
6 Measuring Cosmological Parameters
6.1 Magnitudes and Polar Bear Feet
First solving for the bolometric absolute magnitude of the bear’s foot, we have:
     
L 10W 10W
MB = −2.5 log10 = −2.5 log10 = −2.5 log10 (6.1)
Lx 78.7L 78.7 · 3.82 × 1026 W

Numerically, this is:


MB = 68.7 (6.2)
Now for the apparent magnitude at luminosity distance of d L = 0.5km. We first calculate the flux of the
polar bear foot to be
L 10W
fB = = 3.18 × 10−6 Wm−2 (6.3)
4πd2L 4π (500m)2
So therefore calculating the apparent magnitude, we have:
! !
fB 3.18 × 10−6 Wm−2
m B = −2.5 log10 = −2.5 log10 (6.4)
fx 2.53 × 10−8 Wm−2

Which numerically we calculate to be:


m B = −5.25 (6.5)
If we have a bolometer that can detect the the bear’s foot at a maximum luminosity distance of d L = 0.5km,
then f B (the flux of the bear foot at this distance) defines the maximum flux f max that is capable of being
detected, so solving for the maximum luminosity distance it could detect the sun, we have:
s
L
dmax, sun = = 3.09 × 1015 m (6.6)
4π f max

Repeating this calculation for the supernova, we have:


s s
Lsupernova 4 × 109 L
dmax, supernova = = = 1.96 × 1020 m (6.7)
4π f max 4π f max

6.2 Angular Size and Polar Bear Feet


We know that:
l
dA = (6.8)
δθ
so given l = 0.16m and d A = 0.5km, we can rearrange to solve for the angular size to be:

l 0.16m
δθ = = = 3.2 × 10−4 rad (6.9)
dA 500m

The critical redshift of the benchmark model is zC = 1.6, where d A,max = 5.31 × 1025 m, so:

l
δθmax = = 3 × 10−27 rad (6.10)
d A,max

28
6.3 Maximizing d A in a flat, single-component universe
In a spatially flat, single-component universe, the scale factor is given as (Ryden Eq. 5.39):
 2/(3+3w)
t
a(t) = (6.11)
t0

So we can integrate to obtain the proper distance (Ryden Eq. 5.49):


Z t0 Z t0  −2/(3+3w)
dt t 3(1 + w ) h i
d p ( t0 ) = c =c dt = ct0 1 − (te /t0 )(1+3w)/(3+3w) (6.12)
te a(t) te t0 1 + 3w
a ( t0 )
Furthermore, using that 1 + z = a(te )
= (t0 /te )2/(3+3w) (Ryden Eq. 5.47) to find te , we obtain (Ryden Eq.
5.48):
t0 2 1
te = ( + )
= (6.13)
(1 + z ) 3 3w /2 3(1 + w) H0 (1 + z)3(1+w)/2
Substituting this into (??) we find the current proper distance in terms of z:

c 2 h i
d p ( t0 ) = 1 − (1 + z)−(1+3w)/2 (6.14)
H0 1 + 3w

Further, in the case of a spatially flat universe, we can use Ryden Eq. 6.37 to obtain the current angular
and luminosity distances:

d p ( t0 ) c 2 h i 1
d A ( t0 ) = = 1 − (1 + z)−(1+3w)/2 (6.15)
1+z H0 1 + 3w 1+z

c 2 h i
d L (t0 ) = d p (t0 )(1 + z) = 1 − (1 + z)−(1+3w)/2 (1 + z) (6.16)
H0 1 + 3w
To solve for the critical redshift zC where d A has the maximum value, we take the derivative of (6.15) with
respect to z and set it to 0:
 
dd A c 2 3 + 3w
= −(1 + z)−2 + (1 + z)−(5+3w)/2 = 0 (6.17)
dz H0 1 + 3w 2

The terms in the brackets must vanish, so:


3 + 3w
− (1 + z ) −2 + (1 + z)−(5+3w)/2 = 0 (6.18)
2
Cancelling out a factor of (1 + z)−2 , we get:

3 + 3w
(1 + z)−(1+3w)/2 = 1 (6.19)
2
Now rearranging to solve for zc , we find:

 (1+3w)/2
2
zc = −1 (6.20)
3 + 3w

29
We can now substitute this back into (6.15) to find what the maximum redshift is:

c 2 h i 1
d A,max = 1 − (1 + zc )−(1+3w)/2
H0 1 + 3w 1 + zc
 
(1+3w)/2 −(1+3w)/2 
 
 −(1+3w)/2
c 2 
1 −  2  2
=   (6.21)
H0 1 + 3w  3 + 3w  3 + 3w

So we conclude:
 
 −(1+3w)2 /4  −(1+3w)/2
c 2 2 2
d A,max =  1−  (6.22)
H0 1 + 3w 3 + 3w 3 + 3w

6.4 Difference between relative and absolute magnitude


At small redshift (z  1), the luminosity distance is approximately (Ryden Eq. 6.50, also seen in class):

1 − q0
 
c
dL ≈ z 1+ z . (6.23)
H0 2

By Ryden Eq 6.49, the distance modulus is given by:


!
dL
m − M = 5 log10 + 25. (6.24)
1Mpc

Substituting the first equation into the second, we get:


  
c 1− q0
 H0 z 1 + 2 z
m − M ≈ 5 log10   + 25 (6.25)

1Mpc

From here on out, we will supress the 1Mpc in the denominator for clarity:
!
1 − q0

c
m − M ≈ 5 log10 z 1+ z + 25 (6.26)
H0 2

Using that log( ab) = log( a) + log(b) we get:

1 − q0
   
c
m − M ≈ 5 log10 z + 5 log10 1 + z + 25 (6.27)
H0 2

Using the approximation log10 (1 + x ) ≈ 0.4343 ln(1 + x ) ≈ 0.4343x for small x on the second term, we
get:
1 − q0
 
c
m − M ≈ 5 log10 z + 5(0.4343)( z) + 25 (6.28)
H0 2
Simplifying the numerical terms in the second term, and multiplying by one in the argument of the first
term, we get: !
c 68km s−1 Mpc−1
m − M ≈ 5 log10 z + 1.086(1 − q0 )z + 25 (6.29)
H0 68km s−1 Mpc−1

30
Further application of the log( ab) = log( a) + log(b) rule yields:
! !
c 68km s−1 Mpc−1
m − M ≈ 5 log10 + 5 log10 z + 5 log10 + 1.086(1 − q0 )z + 25
68km s−1 Mpc−1 H0
  (6.30)
Applying the log 1x = − log( x ) rule we get:
! !
c H0
m − M ≈ 5 log10 + 5 log10 z − 5 log10 + 1.086(1 − q0 )z + 25
68km s−1 Mpc−1 68km s−1 Mpc−1
(6.31)
Before we evaluate the first term numerically, we recall the supressed Mpc, so:
! !
300000km s−1 H0
m − M ≈ 5 log10 + 5 log10 z − 5 log10 + 1.086(1 − q0 )z + 25 (6.32)
68km s−1 68km s−1 Mpc−1

We use a calculator to check the value of the first term:


!
H0
m − M ≈ 5(3.6446) + 5 log10 z − 5 log10 + 1.086(1 − q0 )z + 25 (6.33)
68km s−1 Mpc−1

Grouping the numerical terms:


!
H0
m − M ≈ 43.23 − 5 log10 + 5 log10 z + 1.086(1 − q0 )z . (6.34)
68km s−1 Mpc−1

6.5 Surface Brightness


First, we recall the angular diameter distance d A to a standard yardstick to be (Ryden Eq. 6.35):

l Sκ ( r )
dA ≡ = (6.35)
δθ 1+z
Rearranging, we find:
l (1 + z )
δθ = (6.36)
Sκ ( r )
The observed flux is related to the luminosity L and the observed flux f as (Ryden 6.27):
L
f = (6.37)
4πSκ (r )2 (1 + z)2
So, Σ as a function of redshift is:
L
f 4πSκ (r )2 (1+z)2 L 1 1
Σ∝ = = ∝ (6.38)
(δθ )2 l (1+ z ) 2 l 2 4π (1 + z ) 4 (1 + z )4
 
Sκ ( r )

So:
Σ0
Σ= (6.39)
(1 + z )4
for some constant Σ0 . Since the surface brightness Σ only depends on the redshift and not cosmological
parameters, we cannot use it to measure a cosmological parameter q0 .

31
6.6 Quasar Light Flux
The variation time scale at the time light was emitted is related to the variation timescale when it was
observed via:
δt0 = (1 + z)δte (6.40)
So for redshift z = 5.0 and δt0 of 3 days, the variation time scale when emitted is:

δte = 0.5days (6.41)

Rmax for the observed quasar is:


Rmax = c(δte ) = 6.48 × 1012 m (6.42)

From Ryden Figure 6.4, a standard yardstick with redshift z = 5.0 has angular distance d A ≈ 0.3c/H0 in
the Benchmark model, so the angular size is given by:

Rmax 6.48 × 1012 m


δθ = = = 1.58 × 10−44 rad (6.43)
dA 0.3c/H0

6.7 Proper Area of a sphere


The FRW metric (Ryden Eq. 6.22) is:

ds2 = −c2 dt2 + a(t)2 [dr2 + Sk (r )2 dΩ2 ] (6.44)

Expanding out dΩ, this becomes:

ds2 = −c2 dt2 + a(t)2 [dr2 + Sk (r )2 dθ 2 + Sκ (r )2 sin2 θdφ2 ] (6.45)

For a space described by this metric, a surface element dA on a sphere of radius r will be given by:

dA = (Sκ (r )dθ )(Sκ (r ) sin θdφ) = Sκ (r )2 sin θdθdφ (6.46)

Integrating this surface area to find the surface area of this sphere, we have:
ZZ ZZ Z π Z 2π
2 2
A= dA = Sκ (r ) sin θdθdφ = Sκ (r ) sin θdθ dφ = Sκ (r )2 (2)(2π ) = 4πSκ (r )2 (6.47)
0 0

Therefore setting r to be the proper radius r = d p (t0 ), we obtain:

A p (t0 ) = 4πSκ (r )2 . (6.48)

6.8 Total Intensity of Standard Candles


The relation between the observed flux f and the Luminosity L of a distant light source is given by:

L
f = (6.49)
4πSκ (r )2 (1 + z)2

Since we live in a flat universe, Sκ (r ) = r and so:

L
f = (6.50)
4πr2 (1 + z)2

32
In a single-component universe, the proper distance r = d p (t0 ) for w 6= −1/3 is given by (Ryden Eq 5.50):

c 2 h i
r = d p ( t0 ) = 1 − (1 + z)−(1+3w)/2 (6.51)
H0 1 + 3w

So subsituting this into (6.50) we have:

L
f (z) =  h i 2 (6.52)
c 2
4π H0 1+3w 1 − (1 + z)−(1+3w)/2 (1 + z )2

Which we can rewrite as:

L(1 + 3w)2 1 h
−(1+3w)/2
i −2
f (z) = 1 − ( 1 + z ) (6.53)
16π (c/H0 )2 (1 + z)2

which was the desired expression. When w = −1/3, the scale factor in a spatially flat, single component
universe is given by:
 2/(3+3w)  2/(3−1)
t t t
a(t) = = = (6.54)
t0 t0 t0
Therefore, the Hubble constant is given by:
1
ȧ t0 1
H0 = = t0
= (6.55)
a t = t0 t0
t0

The redshift is given by:


 2/(3+3w)
t0 t0
1+z = = (6.56)
te te
Solving for the proper distance r = d p (t0 ) in this universe, we have:
Z t0 Z t0  
dt dt t c
r = d p ( t0 ) = c = ct0 = ct0 ln 0 = ln(1 + z) (6.57)
te a ( t ) te t te H0

Therefore the observed flux in this universe is given by:

L 1
f (z) = (6.58)
4π (c/H0 )2 ln2 (1 + z) (1 + z )2

The number of stars located in the range r to r + dr in the sky per steradian will be given by:

N (r ) = n0 r2 dr (6.59)

So the number of stars in the range z to z + dz per steradian is given by:


 i
2 d c 2 h
−(1+3w)/2
N ( z ) = n0 r 1 − (1 + z ) (6.60)
dz H0 1 + 3w

Taking the derivative:  


2 c −(3+3w)/2
N ( z ) = n0 r (1 + z ) dz (6.61)
H0

33
So, finding the intensity from standard candles in the range z to z + dz, we multiply the earlier result for
f (z) with the above result for N (z):
 
L 2 c −(3+3w)/2
dJ (z) = f (z) N (z) = n0 r (1 + z ) dz (6.62)
4πr2 (1 + z)2 H0

Which after cancelling terms:

n0 L(c/H0 )
dJ (z) = (1 + z)−(7+3w)/2 dz (6.63)

which is the desired result. To find the total intensity J, we integrate over all redshifts:
Z ∞ Z ∞ ∞
n0 L(c/H0 ) −(7+3w)/2 n0 L(c/H0 ) −2
J= dJ (z) = (1 + z ) dz = (1 + z)−(5+3w)/2 (6.64)
0 0 4π 4π (1 + 3w)
0

This gives us the result: (


n0 L(c/H0 ) 1+3w
4π 2 w > − 53
J= (6.65)
∞ w < − 53

To obtain the w = − 31 result, we would have to repeat the analysis using (6.58), but we leave this as an
exercise.
The result above tells us that the total intensity we have in a universe with w < − 53 is infinite! On
some level this makes sense, as the horizon distance is infinite. However, in this universe, we claim the
apparently paradoxical result that the brightness of the night sky is still finite. Why? Because the above
calculation assumes that we see light flux from every single standard candle in the universe; this is simply
NOT the case. There will be standard candles that block the sight of other standard candles to ours, so we
simply do not see the light from every light source in the universe (and hence the above result is actually
misleading, as it does not take into account the fact that light sources block other light sources). The
maximum possible brightness we could have is if stars paved the sky (i.e. every sightline was blocked
eventually by a star). In this scenario, we can repeat the calculation as done in Chapter 2 of Ryden. If a
standard candle of radius R∗ is at a distance r  R∗ , its angular area in steradians is given by:

πR2∗ R2∗
Ω= = (6.66)
4πr (1 + z)2
2 4r (1 + z)2
2

and its measured flux will be:


L∗
f = (6.67)
4πr (1 + z)2
2

so the surface brightness of the star, in watts per square meter is:

f L∗
Σ∗ = = (6.68)
Ω πR2∗

which also gives the surface brightness of the paved sky, which while large, is most certainly finite!

6.9 Expansion Switch


The acceleration equation can be written as:

ä 1 N
− 2
= ∑ Ωi (1 + 3wi ) (6.69)
aH 2 i =1

34
where the sum is taken over the different components of the universe. In the Benchmark model, we have
matter (w = 0), radiation (w = 13 ), and the cosmological constant (w = −1), and so:

ä 1
− = Ω m + Ωr − Ω Λ (6.70)
aH 2 2
Ωm,0 Ωr,0
We have that Ωm = a3
, Ωr = a4
, and ΩΛ = ΩΛ,0 , so:

ä 1 Ωm,0 Ω
− 2
= 3
+ r,0 − ΩΛ,0 (6.71)
aH 2 a a4
Setting ä to find the scale factor a for which the expansion of the universe switched from slowing down
to speeding up, we have:
1 Ωm,0 Ω
0= + r,0 − ΩΛ,0 (6.72)
2 a3 a4
Multiplying by a4 we get:
1
0= Ωm,0 a + Ωr,0 − ΩΛ,0 a4 (6.73)
2
In the Benchmark model, Ωm,0 = 0.31, Ωr,0 = 9.0 × 10−5 , and ΩΛ,0 ≈ 0.69. The radiation term can be
neglected to good approximation, yielding:

1
0 ≈ a( Ωm,0 − ΩΛ,0 a3 ) (6.74)
2
So solving for the positive root of this equation, we get:
s
3
Ωm,0
a= ≈ 0.608 (6.75)
2ΩΛ,0

Which we note is less than the scale factor at matter-lambda equality of amΛ = 0.77.

35
7 Dark Matter
7.1 Dark Matter Candidates
Taking the radius of the halo to be Rhalo ≈ 75kpc and the mass of our galaxy to be Mgal ≈ 9.6 × 1011 M , we
approximate that roughly all of the mass comes from dark matter, hence giving us N = Mgal /10−8 M =
9.6 × 1019 black holes. The volume of our galaxy is given by:

4
V= πR3halo = 5.2 × 1064 m3 (7.1)
3
The number density of these black holes in our galaxy is therefore given by:

N 9.6 × 1019
nBH = = ≈ 1.85 × 10−42 m−3 (7.2)
V 5.2 × 1064 m3
In other words, we can find one black hole per:

1
VBH = = 5.4 × 1041 m3 (7.3)
n BH

So the nearest black hole would (approximately) be a distance of:

VBH = 8.1 × 1013 m


p
3
dBH ≈ (7.4)

away. The mean free path before a black hole comes into a distance 1AU with our sun is given by:

1 1
λBH = = = 7.6 × 1018 m (7.5)
n BH σ n BH π (1AU)2

So combining this with the solar galactic orbital speed of 235km s−1 , we find that the frequency of such
black hole pass-bys are given by:
v
f BH ≈ = 3.1 × 10−14 Hz (7.6)
λ BH

Now we consider MACHOs with mass 10−3 M . The only difference from the previous part is that all
that occurs is N gets scaled by a factor of 10−8 M /10−3 M = 10−5 . This leads to:

dMACHO ≈ 3.8 × 1015 m (7.7)

f MACHO ≈ 3.1 × 10−9 Hz (7.8)

7.2 Draco Galaxy


If we assume that the velocity dispersion is isotropic, then the 3D RMS velocity is equal to the three times
the 1D mean square velocity σr , so:

hv2 i = 3(10.5km s−1 )2 = 3.31 × 108 m s−1 (7.9)

Using the steady state virial theorem, we know that:

1 α GM2
M h v2 i = (7.10)
2 2 rh

36
Which we can rearrange for M:
h v2 ir h
M= (7.11)
αG
Assuming that α = 0.45, we can calculate the mass of the Draco galaxy to be:

3.31 × 108 m s−1 · 120pc


M= == 3.89 × 1033 kg = 2.04 × 107 M (7.12)
0.45 ∗ 6.67 × 10−11 m3 kg−1 s−2

The mass to light ratio is then:


M 2.04 × 107 M M
= ≈ 113 (7.13)
L 1.8 × 105 L L
Some possible errors: The isotropy assumption of the velocity dispersion (on a local scale, the galaxy is
certainly not isotropic). Assuming that the galaxy was in a steady state. Assuming α = 0.45.

7.3 Gravitational Lensing of Earth


The local curvature of spacetime causes the photon to be deflected by angle:

4GM
αEarth = = 2.8 × 10−9 rad (7.14)
c2 R
For a white dwarf and a neutron star, we get:

αdwarf = 4.0 × 10−4 rad (7.15)

αneut = 0.74rad (7.16)

7.4 Halo Mass Density


For a spherically symmetric mass distribution, we can model the density as:
1 d M (r )
ρ (r ) = (7.17)
4πr2 dr
v2 r d M (r ) v2
Since M (r ) = G , = G and so the above becomes:
dr
v2
ρ (r ) = (7.18)
4Gπr2
Note that we have assumed here that v is approximately constant with r. Indeed, we find that v(r ) ≈
230km s−1 out to r = 35kpc for our galaxy, so we are justified as treating it as a constant in our derivation.
Putting in this value for v and G, we get:

1 1
ρ (r ) = 6.3 × 1019 kg m−1 = 2 9.78 × 1011 M Mpc (7.19)
r2 r
If we look at the mass density of the cosmological constant (assuming it to be uniform), we have:

ρλ = ΩΛ ρcrit = 0.7 · 1.28 × 1011 M Mpc−3 ≈ 1011 M Mpc−3 (7.20)

So within our galactic halo (which only extends to ∼ 75kpc), the mass density from the dark matter halo
is evidently is much larger than the cosmological constant; it therefore shouldn’t significaly affect thae
dynamics of our galaxy’s halo.

37
7.5 Cluster Collisions
The number density of galaxies in this half-mass radius is:

N N
n= = 4 3
= 70.7Mpc−3 (7.21)
V 3 πr h

If the typical cross sectional area is Σ ≈ 10−3 Mpc2 , then the mean free path of the Coma cluster before it
hits another galaxy is:
1 1
λ= = = 14.1Mpc (7.22)
Σn (10−3 Mpc2 )(70.7Mpc−3 )

If the velocity dispersion of the Coma cluster is σ ≈ 880km s−1 , assuming isotropy we can obtain the 3D
RMS velocity to be:
hv2 i = 3(880km s−1 )2 = 2.32 × 1012 m2 s−2 (7.23)
p
Then approximating hvi ≈ hv2 i, we get:

hvi ≈ 1.52 × 106 m s−1 = 4.9 × 10−17 Mpc s−1 (7.24)

Hence the average time between collisions is:

λ 14.1Mpc
t= = = 2.9 × 1017 s = 9.2Gyr (7.25)
hvi 4.9 × 10−17 Mpc s−1

The Hubble time is H0−1 ≈ 14Gyr, so t is less than that, but on the same order of magnitude.

7.6 Solar vs. CMB neutrinos


Assuming isotropic emission, neutrinos from the sun (by the time they reach Earth) are spread out over
area:
Ashell = 4πR2 (7.26)
where R = 1AU. Letting Ahuman be the cross-sectional area of a human being, the approximate number
of solar photons that hit us per second is:

Ahuman
rν = rsun (7.27)
Ashell

Modelling the human body as a tube of length Lhuman , the time that a given photon will spend inside the
body is:
L
tν = human (7.28)
c
So the number of neutrinos inside of us at any given moment will be:

Ahuman Lhuman rsun Vhuman


N = rν tν = rsun = (7.29)
Ashell c 4πR2 c

Approximating Vhuman ∼ 0.1m3 , and using R = 1AU = 1.5 × 1011 m, c = 3.0 × 108 m s−1 and rsun =
2 × 1038 neutrinos s−1 (given in the question) we find:

Nsun = 2.4 × 105 neutrinos (7.30)

38
Which gives us our result for the number of solar neutrinos in our body at any given moment. The
number density of neutrinos from the cosmic neutrino background is 3/11 times the number density of
CMB photons (per neutrino flavour), so accounting for the 3 flavours, we get:
 
3 9
nν = 3 nγ = (4.108 × 108 m−3 ) = 3.36 × 108 m−3 (7.31)
11 11

So taking our volume to be roughly V =∼ 0.1m3 , we have approximately:

NCNB = 3.36 × 107 neutrinos (7.32)

cosmic neutrino background neutrinos inside of us at any moment. Hence there are around 100 times
the amount of cosmic neutrino background neutrinos inside of us as there are solar.

39
8 The Cosmic Microwave Background
8.1 Baryon-to-Photon Ratio and Recombination Temperature
We recall the fractional ionization as solved for using the Saha equation:

−1 + 1 + 4S
X= (8.1)
2S
where S is given by:
 3/2  
kT Q
S( T, η ) = 3.84η exp (8.2)
m e c2 kT
where T is the temperature, η is the baryon-to-photon ratio, and Q is the ionization energy. We take
k = 8.62 × 10−5 eV K−1 , Q = 13.6eV, me c2 = 511000eV. For η = 4 × 10−10 and for η = 8 × 10−10 , we get:

X vs. T for eta = 4E-10 X vs. T for eta = 8E-10


1.0 1.0

0.8 0.8

0.6 0.6
X

0.4 0.4

0.2 0.2

0.0 0.0
3000 3200 3400 3600 3800 4000 4200 4400 3000 3200 3400 3600 3800 4000 4200 4400
T (K) T (K)

Figure 6: Plots of fractional ionization X as a function of temperature T (in Kelvin) for baryon-to-photon
ratios η = 4 × 10−10 and η = 8 × 10−10 .

Taking Trec to be when X = 1/2, for η = 4 × 10−10 , we have Trec = 3720K , and for η = 8 × 10−10 ,
we have Trec = 3784K . Doubling the photon-to-baryon ratio has a small effect (only a relative change of
about 1.7%).

8.2 An ionizing photon per baryon


From Problem 2.5, we recall:
 2  
n(h f > E0 ) E0 E0
≈ 0.42 exp − (8.3)
nγ kT kT
So setting n(h f > E0 ) = nbary to have 1 ionizing photon per baryon, and letting E0 = Q, we find:
2
nbary
  
n(h f > Q) Q Q
η= = = 0.42 exp − (8.4)
nγ nγ kT kT

With Q = 13.6eV and η = 6.1 × 10−10 , we can numerically solve the above relation to find:

T = 5823K (8.5)

which is larger than the recombination temperature.

40
8.3 Completely Helium at Recombination
We start with Ryden Eq. 8.28:
!3/2  −3/2 ! !3/2  −3/2
[ m p + m e − m H ] c2
 
nH g mH kT g mH kT QH
= H exp = H exp
n p ne g p gE m p me 2πh̄2 kT g p gE m p me 2πh̄2 kT
(8.6)
The analogous relation for the Helium atom is:
 3/2  −3/2  
n He g He m He kT Q He
= exp (8.7)
ne n He+ ge g He+ me m He+ 2πh̄2 kT

Using that the statistical factor is g He /ge g He+ = 1/4 and that m He ∼ m He+ ∼ 4m p , this becomes:
 −3/2  
n He 1 me kT Q He
= exp (8.8)
ne n He+ 4 2πh̄2 kT
n He+
Defining X = n He+ +n He , we have that:
1−X
n He = n He+ (8.9)
X
and from charge neutrality, n He+ = ne , so:
−3/2
1−X
  
1 me kT Q He
= n He+ exp (8.10)
X 4 2πh̄2 kT

If the Baryonic portion of the universe consists entirely of Helium-4 at the time of recombination, we then
have that:
4n He+
η= (8.11)
Xnγ
Noting the 4 as there are four baryons per Helium nucleus. Therefore using the blackbody number density
of photons (Ryden Eq. 8.23):
 3
X kT
n He+ = 0.2436 η (8.12)
4 h̄c
And plugging this back into (8.10), we have:
3/2
1−X
  
3.84 kT Q He
2
= η exp (8.13)
X 16 m e c2 kT

We now solve for the time of recombination. Setting X = 1/2, we have:


 3/2  
3.84 kTrec Q He
2= η exp (8.14)
16 m e c2 kTrec

We now take η = 6 × 10−10 and Q He = 24.6eV, and substitute in the standard values for the other
constants. We can then umerically solve for Trec to get:

Trec = 6490K (8.15)

41
8.4 Distances to Last Scattering
From Fig 5.9, we can see that at zls = 1090, the propert distance approaches its limiting value of 3.20c/H0 ,
so:
c
dp, ls = 3.20 = 14000Mpc . (8.16)
H0

Finding the luminosity distance is then just multiplying the above by a factor of (1 + zls ):

dL, ls = (1 + zls )dp, ls = 1091dp, ls = 15274Gpc . (8.17)

42
9 Nucleosynthesis and the Early Universe
9.1 Mass fraction of Helium with faster decay
After the proton/neutron freezeout, the ratio of neutrons to protons is approximately:

nn,0 1
f0 = ≈ (9.1)
n p,0 5

Suppose the time delay until nucleosynthesis is t. In this time delay, the neutrons decay to exp(−t/τn ) of
their original amount, and the protons increase by the amount the neutrons decay (as the neutrons decay
into a proton and electron). Therefore, the neutron-to-proton to ratio as a function of delay time is:
nn,0
nn (t) nn,0 exp(−t/τn ) n p,0 exp(− t/τn ) f 0 exp(−t/τn )
f (t) = = = nn,0 =
n p (t) n p,0 + nn,0 (1 − exp(−t/τn )) 1+ n p,0 (1 − exp(− t/τn ))
1 + f 0 (1 − exp(−t/τn ))
(9.2)
The time delay from freezeout until nucleosynthesis is 200s, and we suppose that the neutron decay time
is reduced to τn = 88s, so we can compute the fraction f (200) at the time of nucleosynthesis to be:

f (200) = 0.017 (9.3)

If we assume that all available neutrons are incorporated into Helium, we get the maximal value for the
primordial Helium fraction (as derived in problem 9.4) so:

2 f (200)
Ymax = = 0.033 (9.4)
1 + f (200)

9.2 Mass fraction of Helium with different rest energies


Note that technically, this new difference in the mass fraction would likely cause some difference in the
binding energy of the deuteron. This would affect the temperature Tnuc of nucleosynthesis, as:
 3/2  
kTnuc BD
1 ≈ 6.5η exp (9.5)
m n c2 kTnuc

and therefore the nucleosynthsis time tnuc would also change. For simplicity’s sake, let us assume that BD
remains unchanged. However, the difference in the mass energy will affect the neutron-to-proton ratio at
freezeout, where we find:  
nn,0 Qn
f0 = = exp − (9.6)
n p,0 kTfreeze
so with Qn = 0.129MeV instead of 1.29MeV and kTfreeze = 0.8MeV, we find:

f 0 = 0.85. (9.7)

So, using (9.2) with f 0 = 0.85, t = 200s, τn = 880s, we find:

f (200) = 0.58 (9.8)

therefore the maximum possible mass fraction is given by:

2 f (200)
Ymax = = 0.734 (9.9)
1 + f (200)

43
9.3 Helium Increase in Our Galaxy
With L ≈ 3 × 1010 L , L = 3.83 × 1026 W, and ∆T = 10Gyr, the energy emitted in the form of starlight is:

E = L∆T = 3.62 × 1054 J (9.10)


Assuming that the fusion of Hydrogen into Helium-4 is the only significant energy source, we can find
the number of Helium nuclei produced via simple division:
E 3.62 × 1054 J
NHe = = = 7.17 × 1065 (9.11)
Eper fusion 28.4MeV
The primordial Helium fraction was given by:
ρHe,p MHe,p
Yp = = = 0.24 (9.12)
ρb,p Mb,p
The mass of our galaxy is approximately conserved (there is a very insignificant amount that it decreases
by due to the radiation leaving the galaxy, which we neglect), so Mb,p is equal to the baryonic mass of our
galaxy today. Hence with Mb,p = Mb = 1011 M we have:

MHe,p = 0.24 × 1011 M (9.13)


The current fraction is therefore given by:
MHe MHe,p + NHe mHe
Y= = (9.14)
Mb Mb
so with mHe ≈ 4m p and our previous result for MHe,p , we obtain numerically:
Y = 0.26 (9.15)
So therefore the helium fraction has increased by:
∆Y = Y − Yp = 0.03 (9.16)

9.4 Maximum value for primordial Helium Fraction


By definition, we have:
ρHe
Ymax = (9.17)
ρb
Further, assuming that all neutrons are contained in Helium (as we want the maximal primordial Helium
fraction), we have:
ρHe = 2ρn (9.18)
And the Baryonic density is given by:
ρb = ρn + ρ p (9.19)
Where:
ρn = mn nn , ρp = mpnp (9.20)
So substituting these into (9.17) we have:
nn mn
2ρn 2ρn /ρ p 2 np mp
Ymax = = = (9.21)
ρn + ρ p 1 + ρn /ρ p 1 + nnnp m
mp
n

nn mn
Defining f = np and making the approximation that mp ≈ 1, we conclude:

2f
Ymax ≈ (9.22)
1+ f

44
9.5 Neutrino Detection
The cross-section for the interaction of a neutrino with a proton or neutron is:
 2
−47 2 kT
σw ∼ 10 m . (9.23)
1MeV

So for a typical CNB neutrino with Eν ∼ kTν ∼ 5 × 10−4 eV, we have:

σw ∼ 2.5 × 10−66 m2 (9.24)

Fe-56 has 26 protons, 26 electrons, and 30 neutrons per atom. It has a per-atom weight of:

M = 26m p + 26me + 30mn ≈ 56m p = 56 × 1.67 × 10−27 kg = 9.34 × 10−26 kg (9.25)


So the atomic number density is:

ρ 7900kg m−3
na = = = 8.5 × 1028 m−3 (9.26)
M 9.34 × 10−26 kg
The number density of protons/electrons is therefore:

n p = ne = 26n a = 2.2 × 1030 m−3 (9.27)

and the number density of neutrons is therefore:

nn = 30n a = 2.5 × 1030 m−3 (9.28)

The total number density of sub-atomic particles is therefore:

n ∼ n p + ne + nn = 6.9 × 1030 m−3 (9.29)


though of course this is approximate; really all the subatomic particles are clustered in their respective
atoms. The mean free path of a CNB neutrino in this medium is given by:

1
λ= = 5.8 × 1034 m (9.30)
nσw

Extremely large; neutrinos are very hard to detect!

Optical Depth of Reionized Material (9.5 in first ed.)

We know from ovbservations that the intergalatic medium is currently ionized. Thus, at some point
between trec and t0 , the integalactic medium must have been reionized. In fact detailed measure-
ments of the CMB on large scales place constraints on the amount of reionization (but that isn’t
important for this question). Assume that the baryonic component of the Universe instantaneously
became completely reionized at some time t∗ . For what value of t∗ does the optical depth of the
reionized material: Zt0 Z t0
τ= Γ(t)dt = ne (t)σe cdt (9.31)
t∗ t∗

equal one? For simplicity, assume that the Universe is spatially flat and matter-dominated, and
that the baryonic component of the universe is pure hydrogen. To what redshift z∗ does this alue
of t∗ correspond?

45
As the baryon density scales as ∝ 1
a3
, we have that:

ne,0 σe c
ne (t)σe c = (9.32)
a3
Assuming that the baryonic component of the universe is pure hydrogen and that the universe is charge
neutral, we have:
ne,0 = nbary,0 (9.33)
For a flat, matter-dominated universe, we have scale factor (Ryden 5.5):
 2/3
t
a(t) = (9.34)
t0

so we find:
nbary,0 σe ct20
ne (t)σe c = (9.35)
t2
so carrying out the integral we have:
Z t∗ n 2
bary,0 σe ct0
 
1 1
1=τ= dt = nbary,0 σe ct20 − (9.36)
t0 t2 t0 t∗

So rearranging for t∗ we have:


1
t∗ = 1 1
(9.37)
t0 − nbary,0 σe ct20

In a flat, matter-dominated universe, the age of the universe is given by:

2
t0 = (9.38)
3H0

so we obtain:
1
t∗ = (9.39)
3H0 9H02
2 − 4nbary,0 σe c

Therefore, taking H0 = 68km s−1 Mpc−1 , nbary,0 = 0.25m−3 , c = 3 × 108 m s−1 , and σe = 6.65 × 10−29 m2 ,
we find:
t∗ = 4.2 × 1014 s = 13Myr (9.40)

46
10 Inflation and the Very Early Universe
10.1 Upper limit on Primordial Density
Taking the hint, prior to inflation the Friedmann equation is dominated by the radiation and curvature
term, so:
H2 Ω 1 − Ωr,0
= r,0 + (10.1)
H02 a4 a2
Or writing H = ȧa :
ȧ2 Ω
= r,0 + 1 − Ωr,0 (10.2)
H02 a2
We take our reference time to be at t = t p , so H0 = H p and Ωr,0 = Ω(t p ) and hence:

ȧ2 Ω(t p )
2
= + 1 − Ω(t p ) (10.3)
Hp a2

da
We can write ȧ = dt and integrate both sides to obtain:

a0 da
Z t Z a
Hp dt = q (10.4)
0 0 Ω(t p ) + (1 − Ω(t p )) a02

Where we set a(t = 0) = 0. In order to perform the integral on the LHS, we make the substitution
u = Ω(t p ) + (1 − Ω(t p )) a02 which gives du = 2(1 − Ω(t p )) a0 da0 . It also changes the bouds of integration
to be from Ω(t p ) to Ω(t p ) + (1 − Ω(t p )) a2 . We therefore obtain:
Z Ω(t p )+(1−Ω(t p )) a2
1 1
Hp t = √ du (10.5)
Ω(t p ) 2(1 − Ω(t p )) u

This integral can now be performed to obtain:


!
√ Ω(t p )+(1−Ω(t p ))a2
q 
1 1 q
Hp t = 2 u = Ω(t p ) + (1 − Ω(t p )) a2 − Ω(t p ) (10.6)
2(1 − Ω(t p )) Ω(t p ) 1 − Ω(t p )

We can now solve for a(t):


v
u
u q 2
u (1 − Ω(t p )) H p t + Ω(t p ) − Ω(t p )
u
a(t) = t (10.7)
(1 − Ω(t p ))

In order to find a maximum permissable value of Ω(t p ), we want to see the big crunch exactly at the start
of the inflationary epoch at ti , so:
v
u
u q 2
u (1 − Ω(t p )) H p t + Ω(t p ) − Ω(t p )
u i
a ( ti ) = t =0 (10.8)
(1 − Ω(t p ))

From this we obtain: q q


(1 − Ω(t p )) H p ti + Ω(t p ) = ± Ω(t p ) (10.9)

47
q
If ti = 0 we have that the LHS equals + Ω(t p ), but since we want ti > 0, we solve for the negative
solution. this yields: q
−2 Ω ( t p )
H p ti = (10.10)
1 − Ω(t p )
1 1
From the previous question, we know that H = 2t in a radiation-dominated universe, so H p = 2t p and so:
q
ti 2 Ω(t p )
= (10.11)
2t p Ω(t p ) − 1
ti
For conveninece, let us define α = 4t p :
q
Ω(t p )
α= (10.12)
Ω(t p ) − 1
Rearranging, we find a quadratic equation in Ω(t p ):

2
16t2p
Ω ( t p ) − (2 + )Ω(t p ) + 1 = 0 (10.13)
t2i

This has solutions: r


16t2p 16t2p 2
(2 + t2i
)± (2 + t2i
) −4
Ω(t p ) = (10.14)
2
Since Ω(t p ) > 1, we reject the negative solution above. So, the maximum possible Ω(t p ) is given by:
r
16t2p 64t2p 256t4p
(2 + t2i
)+ t2i
+ t4i
Ω(t p ) = (10.15)
2
tp
Since t p  ti , we can neglect all powers of ti in the above expression that are higher than linear order.
This yields:
tp
Ω(t p ) ≈ 1 + 4 (10.16)
ti

For ti = 10−36 s, we have:


Ω(t p ) = 1 + 2 × 10−7 (10.17)

For ti = 10−26 s, we have:


Ω(t p ) = 1 + 2 × 10−17 (10.18)

10.2 Solving the Monopole Problem


If monopoles formed at the GUT time with one monopole per horizon of mass m M = mGUT , then the
energy density of these monopoles would be given by:

m M c2
e M (tGUT ) ≈ = 10106 eV m−3 (10.19)
(2ctGUT )

48
where we take m M c2 ∼ EGUT ∼ 1012 TeV and tGUT = 10−36 s. Therefore the density parameter of the
monopoles at this time is given by:

e M (tGUT )
Ω M (tGUT ) = ≈ 1096 (10.20)
ec,0

where we take ec,0 = 5 × 109 eV m−3 . For a radiation-only universe, the scale factor goes as (Ryden Eq.
5.60):
 1/2
t
a(t) = (10.21)
t0
So at the GUT time:  1/2
tGUT
a(tGUT ) = (10.22)
t0
Taking t0 ∼ 13.8Gyr = 4.35 × 1017 s and tGUT as before, we find:

a(tGUT ) = 1.5 × 10−27 . (10.23)

Since the magnetic monopole density parameter should scale as a13 with time (like regular matter), if
inflation did not happen, the density of monopoles today would be:

Ω M (t0 ) = Ω M (tGUT ) a(tGUT )3 = 3.4 × 1015 (10.24)

This is off from the observational limits by:

Ω M,0,observed 10−6
= = 2.9 × 10−22 . (10.25)
Ω M ( t0 ) 3.4 × 1015

To account for this, we must have had N e-folds of inflation, leading the scale factor is actually e N larger
than what we calculated above. Hence:
1
2.9 × 10−22 = 3N (10.26)
e
which we can solve for N to obtain:
N ≈ 17 (10.27)

10.3 False Vacuum


The Universe dominated by the false vacuum has the same structure as one with dominated by a cosmo-
logical constant. In such a universe, the Hubble parameter s given by Ryden 5.72:
 1/2
8πGeΛ
Hi = = 1.83 × 10−18 s−1 = 0.83H0 (10.28)
3c2

where we take H0 is the Hubble constant in our universe. We assume that at the false vacuum decay time
that the false vacuum decays into blackbody photons, so we can therefore use the blackbody radiation
temperature to solve for what the temperature of the Universe would be at this time:
r
4 eΛ
T= = 29K (10.29)
α

To find the energy density of matter at this time, we first compute the scale factor at this time; this is given
by Ryden 5.73 to be:
a(t f ) = a(t0 )e Hi (t f −t0 ) = e Hi (t f −t0 ) (10.30)

49
where we take a(t0 ) = 1 by convention. So with Hi as above, t f = 50t0 and t0 = 13.7Gyr, we find:

a(t f ) = 6.7 × 1016 (10.31)

So the energy density of matter at this time would be:

em (t f ) = a(t f )−3 em,0 = 5 × 10−48 MeV m−3 (10.32)

To find when the universe is again dominated by matter, we wish to find the time when:

e m = er . (10.33)

Further, we know that:


em,0 er,0
em ( a ) = , er ( a ) = (10.34)
a3 a4
First solving for what er,0 would be, we have:

er,0 = er (t f ) a(t f )4 = 6.7 × 1070 MeV m−3 (10.35)

Combining (10.33) and (10.34) we find:


em,0 e
3
= 4r,0 (10.36)
arm arm
Hence:
er,0
arm = (10.37)
em,0
 1/2
t
Further in a radiation dominated universe we know we have a(t) = t0 (Ryden Eq. 5.60), so:

!2
er,0
trm = t0 (10.38)
em,0

Numerically we obtain:
trm = 2.7 × 10136 Gyr (10.39)

Gamow’s CMB Prediction (10.3 in first ed.)

A fascinating bit of cosmological history is that of George Gamow’s prediction of the Cosmic
Microwave Background in 1948. (Unfortunately, his prediction was premature; by the time the
CMB was actually discovered in the 1960’s, his prediction had fallen into obscurity.) Let’s see if
you can reproduce Gamow’s line of argument. Gamow knew that nucleosynthesis must have taken
place at a temperature Tnuc ' 109 K, and that the age of the Universe is currently t0 ' 10Gyr.
Assume that the Universe is flat and contains only radiation. With these assumptions, what was
the energy density e at the time of nucleosynthesis? What was the Hubble parameter H at the
time of nucleosynthesis? What was the time tnuc at which nucleosynthesis took place? What is the
current temperature T0 of the radiation filling the Universe today? If the Universe switched from
being radiation-dominated to being matter-dominated at a redshift zrm > 0, will this increase or
decrease T0 for fixed values of Tnuc and t0 ? Explain your answer.

If the Universe only contains radiation, the energy density is given by the black body energy density
formula:
eγ ( T ) = αT 4 (10.40)

50
So with α = 7.566 × 10−16 J m−3 K−4 and Tnuc = 109 K we have:

4
enuc = αTnuc = 7.566 × 1020 J m−3 (10.41)

Furthermore, Ryden Eq. 5.63 gives the energy density in a radiation-only universe as a function of time to
be: ! −2
Ep t
e(t) = 0.030 3 (10.42)
lp tp
We can invert the above formula to find tnuc :
s
Ep tp
tnuc = 0.030 (10.43)
l 3p
p
e(tnuc )

which numerically gives:


tnuc = 230s (10.44)
Next, in a radiation dominated universe, we have (from Ryden Eq. 5.60):
 1/2
t
a(t) = (10.45)
t0

So the Hubble parameter is given by:


ȧ 1
H= = (10.46)
a 2t
So at nucleosynthesis:
Hnuc = 2.17 × 10−3 s−1 (10.47)
1
The energy density of radiation scales as a4
, so we can use this to find the energy density of the universe
today:
 2
tnuc
er,0 = enuc a4nuc = enuc (10.48)
t0
so we can find the current temperature (which is still described by a black body) as:
 2
tnuc
αT04 = αTnuc
4
(10.49)
t0
 1/2
tnuc
T0 = Tnuc (10.50)
t0
Numerically we find:
T0 = 27K (10.51)
Now, we suppose that the Universe switched from being radiation-dominated to matter-dominated at
some redshift zrm > 0. This would not change the energy density at the time of nucleosynthesis, but the
scale factor would now grow as:
 2/3
t
a(t) = (10.52)
t0
instead of the previous ∝ t1/2 . Hence the energy density of the radiation would go down more rapidly than
in the radiation-dominated-for-all times Universe, and hence the temperature T0 today would decrease .

51
11 Structure Formation: Gravitational Instability
11.1 Density Fluctuations in a Flat, Matter-Dominated, Contracting Universe
We first note that:

H=<0 (11.1)
a
in a contracting universe. Now, starting with Ryden Eq. 11.49, we have:
3
δ̈ + 2H δ̇ − Ωm H 2 δ = 0 (11.2)
2
In a contracting matter dominated universe, Ωm = 1 and H = − 3t
2
(note the negative sign for contraction!
 2/3
t
This can be obtained from a(t) = t0 ), so:

4 2
δ̈ − δ̇ − 2 δ = 0 (11.3)
3t 3t
Guessing a power law δ(t) = tr , we find:
4 r −1 2
r ( r − 1 ) t r −2 − rt − 2 tr = 0 (11.4)
3t 3t
Dividing both sides by tr−2 we obtain a quadratic equation for r:
7 2
r2 − r − = 0 (11.5)
3 3
Using the quadratic formula, we find the solutions:

7± 73
r= (11.6)
6
So we therefore find:
√ √
7+ 73 7− 73
δ(t) = At 6 + Bt 6 (11.7)
Where A, B are constants determined by initial conditions. The second term vanishes for large t, so:

7+ 73
δ(t) ≈ At 6 (11.8)

11.2 Density Fluctuations in a Nearly Empty, Negatively Curved, Expanding Uni-


verse
We again start with Ryden Eq. 11.49:
3
δ̈ + 2H δ̇ − Ωm H 2 δ = 0 (11.9)
2
Since Ωm  1, the last term can be neglected:

δ̈ + 2H δ̇ = 0 (11.10)
1 t
In an empty expanding universe, we have H = t (as a(t) = t0 ) so:

2
δ̈ + δ̇ = 0 (11.11)
t

52
Again guessing a power law δ(t) = tr , we find:

2
r (r − 1)tr−2 + rtr−1 = 0 (11.12)
t
Dividing both sides by tr−2 we obtain a quadratic equation for r:

r2 + r = 0 (11.13)

This has solutions:


r = 0, r = −1 (11.14)
So we therefore find:
δ(t) = At0 + Bt−1 (11.15)

At long times, the latter terms vanishes, so:

δ(t) ≈ A . (11.16)

That is, at long times the fluctuations are constant.

11.3 Photon-Baryon Fluid


For the photon-baryon fluid, we have:
  −1
dP dP da dP de
= = (11.17)
de da de da da

Since P = Pγ = 13 eγ = 13 eγ,0 a−4 , we find:

dP 4
= − eγ,0 a−5 (11.18)
da 3

Furthermore, since e = eγ + ebary = eγ,0 a−4 + ebary,0 a−3 we find:

de
= −4eγ,0 a−5 − 3ebary,0 a−4 (11.19)
da
Therefore:  
dP − 43 eγ,0 a−5 1 1
= =  (11.20)
de −4eγ,0 a−5 − 3ebary,0 a−4 3 1 + 3 ebary,0 a
4 eγ,0

. Now, we can find the sound speed to be:


v  
 1/2 u
dP u1  1
u
cs = c = ct  (11.21)
de 3 1 + 3 ebary,0 a
4 eγ,0

In Ryden Eq. 11.26, the Jeans length of the photon-baryon fluid neglecting the contribution from the
baryons was estimated to be:
c
λ J,γ ≈ 3 ≈ 0.66Mpc (11.22)
H (zdec )

53
We can replace c with cs in the above expression to determine λ J accounting for the baryons. Before this,
we briefly resolve some remaining parameters in our expression for cs . zdec = 1090, so:
1
adec = = 9.17 × 10−4 (11.23)
1 + zdec
Furthermore, from the benchmark model we find the energy density ratio of:
ebary,0 Ωbary, 0 0.048
= = = 897 (11.24)
eγ,0 Ωγ,0 5.35 × 10−5
ebary,0
Plugging these values of adec , eγ,0 into (11.21), we find:

cs = 0.45c (11.25)

compared to the cs = √c = 0.577c prediction neglecting baryons. Hence the Jeans length being propor-
3
tional to cs gets scaled down by this ratio:
0.45c
λ J,bary+γ = λ J,γ = 0.51Mpc (11.26)
0.577c
So we were off by:
λ J,γ − λ J,bary+γ = 0.15Mpc (11.27)

11.4 Milky Way Gravitational Collapse


For any circular orbit (as we can assume for a spherical dark halo), the orbital speed is related to Rhalo
and Mgal by (Ryden Eq. 7.9):
GMgal
v2 = (11.28)
Rhalo
Further assuming a uniform distribution of mass throughout the galaxy, we have:
Mgal Mgal
ρ= = 4 3
(11.29)
V 3 πRhalo

Combining (11.28) and (11.29), we obtain:


Rhalo v2
G v2
ρ= 4 3
= 4 2
(11.30)
3 πR halo 3 πGRhalo

Substituting this into tmin ≈ √1 we find:



r
1 3 Rhalo
tmin ≈ r = (11.31)
v2 4π v
G4 2
3 πGRhalo

So we therefore conclude:
Rhalo
tmin ∼ (11.32)
v
For our own Galaxy, we take Rhalo ≈ 75kpc and v = 235km s−1 to find:

tmin ≈ 9.86 × 1015 s = 312Myr (11.33)

54
tmin defines the maximum possible redshift in which we could see galaxies with comparable to v and
Rmin ; in other words we find z(tmin ). Neglecting the early radiation period, we can take our universe to
be dominated by matter and the cosmological constant. For this Universe, the Friedmann equation has
the analytic solution (Ryden Eq. 5.101):
 s 
 3/2  3
2 a a
H0 t = p ln  + 1+  (11.34)
3 1 − Ωm,0 amΛ amΛ

With the Benchmark models of H0 = 68km s−1 Mpc−1 , Ωm,0 = 0.31, and amΛ = 0.77, and the pervious
result that t = tmin = 312Myr, we can numerically solve for a in the above relation. Doing so, we find:

amin = 0.069 (11.35)

From which we obtain the maximum redshift:

1
zmax = − 1 = 13.5 (11.36)
amin

11.5 Coma Cluster Gravitational Collapse


We first note (from Ryden 7.37) that: D E
v2 r h
MComa = (11.37)
αG
Taking again ρ to be uniform, we find:
D E
h v2 i r h 3 v2
MComa αG
ρ= = = (11.38)
V 4 3
3 πr h
4παGr2h

So using that tmin ≈ √1 we find:



r
1 4πα rh r
tmin ≈ r = q ∼ q h (11.39)
3 h v2 i 3 v2 v2
G 4παGr2
h

So numerically we find:
tmin ≈ 3.05 × 1016 s = 967Myr (11.40)

11.6 Mean Square Mass Fluctuation and Standard Deviation of Density Field
The standard deviation of the density field for a Gaussian field can be computed (in the case of a Gaussian
field) as (Ryden Eq. 11.67):
Z ∞ Z k max
V V
σδ2 = 2
P(k)k dk = P(k)k2 dk (11.41)
2π 0 2π 0

Meanwhile the mean square mass fluctuation is computed as:


* 2 + * !2 + Z ∞
" #2 Z kmax
" #2
δM M − h Mi V 3j1 (kr ) 2 V 3j1 (kr )
= = P(k) k dk = P(k) k2 dk
M h Mi 2π 0 kr 2π 0 kr
(11.42)

55
So the claim is proven is we can show for M < Mmin , or for r < rmin that:

Z k max Z kmax
" #2
2 3j1 (kr )
P(k)k dk = P(k) k2 dk (11.43)
0 0 kr

Incoming is a handwavey argument that I don’t actually think is the full answer. For r < rmin = k2π
max
, kr
will be small over the domain of integration so we may Taylor expand j1 (kr ). Doing so, we find:
 
(kr )3 (kr )2
(kr − 6 ) − kr 1 − 2 (kr )3
sin(kr ) − kr cos(kr ) kr
j1 (kr ) = ≈ = 3 2 = (11.44)
(kr )2 (kr )2 (kr ) 3

Therefore: " #2
Z kmax Z kmax  2 Z kmax
3j1 (kr ) 3 kr
P(k) k2 dk ≈ P(k) k2 dk = P(k)k2 dk (11.45)
0 kr 0 kr 3 0

Which is exactly what we wished to show.

56
12 Structure Formation: Baryons and Photons
12.1 Galaxies more luminous than L
We start with the Schechter luminosity function which gives the number density Φ( L)dL in the range
L → L + dL:  α  
L L dL
Φ( L)dL = Φ∗ exp − (12.1)
L∗ L∗ L∗
The number density of galaxies more luminous than L is given by integrating from L to ∞:
Z ∞ Z ∞
!α !
0 0 ∗ L0 L0 dL0
n≥ L = Φ( L )dL = Φ exp − ∗ (12.2)
L L L∗ L L∗

L0 dL0
Now substituting t = L∗ which gives dt = L∗ we find:
Z ∞
n≥ L = Φ∗ tα e−t dt (12.3)
L/L∗

Furthermore, we can recognize the RHS as an incomplete gamma function, yielding:

L
n≥ L = Φ∗ Γ(α + 1, ) (12.4)
L∗
In the limit L → 0, we just have the regular gamma function:
Z ∞
n≥0 = ntot = Φ∗ tα e−t dt = Φ∗ Γ(α + 1) (12.5)
0

But with α = −1, the integral:


Z ∞ −t
e
ntot = dt = ∞ (12.6)
0t
is seen to diverge (at the lower limit). The physical solution is as follows; in this limit, the number density
of galaxies diverges, but the luminosity density remains finite (so the average luminosity per galaxy is
zero); this can be realized via the following calculation:
Z ∞ Z ∞
! −1 ! Z ∞
!
0 0 0 ∗ L0 L0 dL0 ∗ L0
Ψtot = L Φ( L )dL = LΦ exp − ∗ =Φ exp − ∗ dL0 = Φ∗ L∗ . (12.7)
0 0 L∗ L L∗ 0 L

12.2 Total Luminosity Density


We again tart with the Schechter luminosity function:
 α  
∗ L L dL
Φ( L)dL = Φ exp − ∗ (12.8)
L∗ L L∗
We can now integrate LΦ( L)dL from 0 to ∞ to obtain the total luminosity density:
Z ∞ Z ∞
!α ! Z ∞
! α +1 !
0 0 0 ∗ L0 L0 dL0 ∗ L0 L0
Ψ= Φ( L )dL = LΦ exp − ∗ = Φ exp − ∗ dL0 (12.9)
0 0 L∗ L L∗ 0 L∗ L

L0
Again making the substitution t = L∗ we find:
Z ∞ Z ∞
Ψ= Φ∗ L∗ tα+1 e−t dt = Φ∗ L∗ tα+1 e−t dt (12.10)
0 0

57
We recognize the rightmost expression as a gamma function, which yields:

Ψ = Φ ∗ L ∗ Γ ( α + 2) (12.11)

Now, observing that Γ(−1 + 2) = Γ(1) = 1, so we find for the V band that:

ΨV = Φ ∗ LV

Γ(1) = 1 × 108 L ,V Mpc
−3
(12.12)

12.3 Do structures of 1017 M exist today?


From Ryden Eq. 12.38, the total amount of mass inside the last scattering surface is given by:

Mtot ≈ 4.3 × 1023 M (12.13)

Dividing this by M = 1017 M , we find:

Mtot
N= = 4.3 × 106 regions (12.14)
M
The very first M = 1017 M structure to collapse is the one region out of 4.3 million that had the highest
overdensity at the time of radiation-matter equality, with probability:

1
P= = 2.3 × 10−7 . (12.15)
N
This is equivalent to a 5.04σ deviation in a Gaussian distribution. Since σ = δM/M = 0.12, then we can
compute the redshift of collapse to be:

1 + zcoll = 5.04σ = 0.6048 (12.16)

i.e. the first such object has not begin to collapse (as zcoll < 0) and hence we do not expect to see
gravitationally collapsed structures with mass M = 1017 M today.

12.4 Ripping Apart Galaxies


An object is ripped apart when its energy density em becomes less than the phantom energy density e p .
Hence we can set these two densities to be equal:

em = e p . (12.17)

We know that e p ∝ a−3(1+w p ) , and in particular:

e p = Ω p,0 ec a−3(1+w p ) . (12.18)

So solving for the rip-apart scale factor a we find:


!−1/(3+3w p )
em
a= (12.19)
Ω p ec

and for w p = −1.1 this becomes:


!10/3
em
a= . (12.20)
Ω p ec

58
For the milky way galaxy, we have Mgal = 9.6 × 1011 M and Rgal = 75kpc. Assuming a uniform mass
density and using that ρc = 8.7 × 10−27 kg m−3 , we find:
egal ρgal Mgal
= = 4 3
= 4.2 × 103 (12.21)
ec ρc 3 πRgal ρc

Therefore the scale factor for which the Milky way galaxy would be ripped apart can be solved to be:

agal = 4.0 × 1012 (12.22)

We can do the same for the sun, with mass M and radius R = 7 × 108 m to find:
esun ρsun M
= = 4 3
= 1.6 × 1029 (12.23)
ec ρc 3 πR ρc
so the scale factor for which the sun would be ripped apart would be:

asun = 7.2 × 1097 (12.24)

From Problem 5.5, we know the time between the present time t0 and the Big rip trip was found for
H0 = 68km s−1 Mpc−1 , w p = −1.1 and Ωm,0 = 0.3 to be:

trip − t0 = 115.5Gyr (12.25)

with t0 = 13.8Gyr, we can find trip to be:

trip = 115.5Gyr − 13.8Gyr = 101.7Gyr (12.26)

Further, we can use the intermediate result from problem 5.5 (Eq. (5.43)) that:
Z trip Z ∞ Z ∞
1
H0 dt ≈ a3(1+w p )/2 (1 − Ωm,0 )−1/2 da = p a3(1+w p )/2−1 da (12.27)
t0 1 Ω p,0 1

Now, we can replace t0 with tgal (the rip-apart time for the Milky Way Galaxy), and a(t0 ) = 1 with
a(tgal ) = agal in the above equation, and carry out the integral to get:

1 2
H0 (trip − tgal ) = p a3(1+w p )/2 (12.28)
Ω p,0 3(1 + w p )
agal

The term at infinity vanishes (as the exponent of a is negative), and we therefore find:
1 2 3(1+w p )/2
tgal = trip − a (12.29)
Ω p,0 3 1 + w p gal
p
H0

Numerically, this evaluates to:

tgal = 101.7Gyr − 0.46Gyr = 101.wGyr (12.30)

So we find that the Milky way will be ripped apart about half a gigayear before the big rip. We can do the
same with tsun and a(tsun ) = asun to find:

tsun = 101.7Gyr − 7.5 × 10−14 Gyr ≈ 101.7Gyr (12.31)

In other words, the sun will not be ripped apart until the big rip.

59

You might also like