Download as pdf or txt
Download as pdf or txt
You are on page 1of 89

Cosmology

Cosmological Models
(mainly relativistic Cosmology)
[References]
Objects in the universe are essentially electrically neutral, such that gravity is the
force driving the dynamics of the universe. Therefore the basic theory is general
relativity theory, which in the first instance expresses our local experience.
µν
One has to make other assumption: T universe, boundary conditions, etc.

There are a number of global properties we can assign to the universe: radius,
curvature, mean radiation density and mean mass density as well as dynamical
properties like the expansion or the fluctuations in the microwave radiation.

Basic Postulates:

One of the basic observations is that velocities of matter systems on astronomical


vicinities are small, which together with the validity of the equivalence principle (all
freely falling bodies move in the same way) suggest the validity of a
hydrodynamic model (streaming fluid) and hence to the assumption:

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 301
Cosmology

Ê Weyl’s Postulate:
The world lines form a bundle of geodesic lines, which diverge from a point (finite
of infinite) in the past or in the future.

This means that geodesic lines meet in not more than one single singular point.

Remark: it should be stressed that this is a strong idealization and a special


assumption about the true distribution and movement of matter. At the same time
it is natural in the sense of the equivalence principle: matter moves independent of
individual properties universal under the influence of gravitation.

Consequence: there exist a bundle of hypersurfaces orthogonal to the geodesics.

Which means: matter defines e natural coordinate system, the co-moving rest
system of matter:

l xi = constant along world lines

l t = constant on hypersurfaces orthogonal to the world lines

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 302
Cosmology

Consequently: a co-moving observer measures t as proper time

Therefore the line element in the co-moving system takes the form:

ds2 = (c dt)2 + gik dxi dxk ; i, k = 1, 2, 3

with t the cosmic time. The latter defines a universal simultaneity, which is actually
necessary to have a meaningful application of GRT to the universe.

Note: the existence of a cosmic time is preconditioned by the way the universe is
populated with matter (galaxies etc.).

Exercise: discuss whether the postulate of a universal time is in contradiction to


the special relativity principle?

Ë Cosmological Principle:
A co-moving observer xi = constant is not able to distinguish his position nor any
direction.

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 303
Cosmology

Which means: the subspaces t = constant are homogeneous and isotropic


(thereby the homogeneity is a consequence of isotropy, see below).

Notice: this assumption is very special as it implies that for t = constant density
ρ(x) and pressure p(x) are constant in the entire universe.

Consequence: the subspaces t = constant are geometrically similar in the sense


that the spatial metric has the form:

gik = −S 2(t) hik ; S 2(t) > 0 ,

and hik is homogeneous, isotropic and time-independent.

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 304
Cosmology

In fact hik is the metric of a 3D space of constant curvature (any point is center):

 2
i
xidx
i
hik dx dx k
{zdx} + a2 − xi xi
= δ|ik dx i k

dr2 +r2 dΩ2

dr2  
= + r dθ + sin θ dϕ
2 2 2 2
1 − r2/a2

in spherical coordinates. For a → ∞ we have flat 3D-space. Note that


(xidxi)2 = r2 (dr)2 and

2
r2 a2
dr 2
(dr)
dr 2
+ dr 2
= 2 =
2
a −r 2 a −r 2 1 − r2/a2

with the above form as a result.

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 305
Cosmology

>0



1/a2 is the space curvature =  =0


 <0

For finite a, the transformation of variables:

k = a2/α2 ; α = |a| ; r/α → r ; α S (t) → S (t)

leads to the well known Robetson-Walker (RW) line element of a homogeneous,


isotropic cosmological model:

dr2
n  o
ds = (c dt) − S (t)
2 2 2
1−kr2
+r 2
dθ + sin θ dϕ
2 2 2

The case k = 0 corresponding to 1/a2 = 0, which remains unchanged.

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 306
Cosmology

with


 1 is the positive spherical closed
k=

0 space zero flat infinite


 −1 curvature negative hyperbolic open

whereby the transformation r → r/(1 + k r2/4) leads back to the standard form.

Key points:

r hypersurfaces t = constant are 3D spaces of constant curvature

r time development is a conformal mapping of he 3D geometries


with scale factor S (t).

r distances between two points are proportional to S (t) everywhere.


If S (t) is increasing we have a general expansion.

r the spatial structure and the motion of galaxies cause each other.

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 307
Cosmology

We mention here another interesting fact: Infeld & Schild have shown that every
homogeneous isotropic model is conformally flat6 (Minkowskian):


ds = e
2 2ζ(x) gµν dxµdxν

The transformation (t, r) → (τ, ρ) which maps the RW line element to manifestly
conformal form reads:

6 The mapping

ds2 → ds2 = α2 (x) ds2 ; α2 > 0

is a conformal transformation (specifically a dilatation)

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 308
Cosmology

Zct 1/2Z(cτ+ρ) 1/2Z(cτ−ρ)


dz dx dx
= +
S (z) 1 + k x2 1 + k x2
0 0 0

Zr 1/2Z(cτ+ρ) 1/2Z(cτ−ρ)
du dx dx
√ = − .
1 − k u2 1+kx 2 1+kx 2
0 0 0

One gets

2
( )
dr 
ds 2
= (c dt) − S (t)2 2
+ r 2
dθ 2
+ sin2
θ dϕ 2
1 − kr2
n h  io
= e (c dτ) − dρ + ρ dθ + sin θ dϕ
2ζ 2 2 2 2 2 2

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 309
Cosmology

with conformal function

S 2(t)
e =

.
[1 + k/4 (cτ + ρ) ][1 + k/4 (cτ − ρ) ]
2 2

This result is important because Electrodynamics is conformally invariant.


Maxwell’s equations thus apply without modification to homogeneous, isotropic
world models.

Exercise: Prove the conformal invariance of Maxwell’s equations.

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 310
Cosmology

Observable Quantities
For the physical interpretation the influence of the RW-metric

2
( )
dr 
ds2 = (c dt)2 − S 2(t) + r 2
dθ 2
+ sin2
θ dϕ 2
1 − kr2

on observable quantities must be discussed.

Exercise: Show that in the co-moving spatial coordinates (r, θ, ϕ)

2
dr
d`2 = + r 2
dΩ2
1 − kr2

is invariant under the following transformations:

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 311
Cosmology

0
a) Rigid rotations: xi = Rik xk ; RRT = 1
n√ √
i0 ~x·~a
  o
b) Quasi-translations: x = x + a 2 i
1−kr − 1− 1−ka2 i
a2


where a is an arbitrary vector with 1 − k a2 real, and a2 = aiai. This
i

transformation maps the origin xi = 0 into the point ai.

For later use we note that for


1
 

1−k r2
0 0 
gik = S 2(t) hik ; hik =  r2
 
0 0 
0 0 r2 sin2 θ
 

the determinant of the spatial metric is given by

1
det gik = S (t) det hik ; det hik =
6
r 4
sin2
θ.
1 − k r2

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 312
Cosmology

Spectral Shift (Red Shift)


For radial light propagation ds = 0 (null geodesics) and hence

c dt
S (t) =± √ dr
1−k r2

A light source Q1 which in its rest frame (t1, r1) has period ∆t1. An observer in its
rest system (t0, r0) has a reference source Q0 of period ∆t1 and he observes a
period ∆t0 from Q1. Assuming ∆t0, ∆t1  |t0 − t1| we denote by λ = c∆t the wave
length and by ν = 1/∆t the frequency.

One then defines the spectral ratio

ν1
ζ =1+z ν0 = 1 + ∆λ
λ1 ; ∆λ = c (∆t0 − ∆t1 )

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 313
Cosmology

z is the red shift:

z>0
(
∆λ red shift
z= λ1 z<0 blue shift

For the RW-metric we obtain:

Rr1 Rt0
1
c
√ dr = dt
S (t)
1−k r2
r0 =0 t1

independent of t!

The r integral yields

Zr k= 1
 
arcsin r ;
dr0 r3

 

= k= 0 = r + k · + O(r5)
 
r ;
 

6

02
1−kr k = −1

 arcsinh r 
;
 

0

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 314
Cosmology

The time independence implies


Zt0 0 +∆t0
tZ
dt dt
t
t0 + ∆t0
=
S (t) S (t)
t1 t1 +∆t1

0 +∆t0
tZ 1 +∆t1
tZ
t0 r dt dt
= −
S (t) S (t)
t1 t1
t1 + ∆t1 0 +∆t0 1 +∆t1
t1 Zt0 tZ tZ
dt dt dt
= + −
r0 = 0 r1 S (t) S (t) S (t)
t1 t0 t1

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 315
Cosmology

Notice: t1 < t0, t1 emission time, t0 receiving time, as ∆t0, ∆t1  |t0 − t1| (clock
period small with respect to transmission time) we must have:

∆t0 ∆t1 ∆t0 S (t0 )


S (t0 ) = S (t1 ) hence ζ =1+z= ∆t1 = S (t1 )

Result: the time dependent scale factor S (t) implies a red shift
S (t0 )
z= S (t1 ) −1

Expansion: S (t0) > S (t1) growing  z>0 red shift


stationary S (t) = constant ./ z = 0 no shift
Contraction: S (t0) < S (t1) decreasing  z<0 blue shift

[Link to Lect.: 13]

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 316
Cosmology

Observational fact:

Ê for very distant galaxies (d ≥ 1 Gpc) one indeed observes a systematic red shift
(expansion, Hubble’s law)

Ë for nearer objects (d ∼< 1 Gpc) z ' 0 which tells us that S (t) is very slowly varying.

One therefore can make a number of approximations, by expanding in τ = t0 − t1.

First we define:

Ṡ (t0 )
H0 = H(t0) = S (t0 ) Hubble constant
S̈ (t0 ) S (t0 )
q0 = q(t0) = − 2 deceleration parameter
(Ṡ (t0))

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 317
Cosmology

and then write

τ 2 2
( )
S (t1) = S (t0) 1 − τ H0 − H0 q0 − · · ·
2
τ2
( )
S −1(t1) = S −1(t0) 1 + τ H0 + H02 (2 + q0) + · · · .
2

Thereby τ is determined by the distance to the source. For the red shift we then
have

τ2
z = τ H0 + 2 H02 (2 + q0) + · · ·

Exercise: show that

1 dz Ṡ (t0) − Ṡ (t1)
= ' −q0 H0 .
z dt0 S (t0) − S (t1)

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 318
Cosmology

Measuring Distances

Distance measurements on cosmological scales are quite problematic and


correspondingly there are quite a number of definitions of distances.

The coordinate distance between an event (t1, r1) and the observer (t0, r0 = 0) is
r1 .

The proper distance D1 can be defined by setting dt = 0:

Zr1 k=1

arcsin r1 ;
dr


= S (t0) = S (t0) ·  k=0

D1 r1 ;


1 − k r2 k = −1

 arcsinh r ;

0 1
 
3
 r1 
= S (t0)  + +
 
r k · · ·
 
1
 6


c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 319
Cosmology

Zt0
S (t0)
D1 = c dt
S (t)
t1

Z0 02
τ
!
= c dτ0 1 − τ0 H0 + H02 (2 + q0) + · · ·
2
−τ=t0 −t1

τ τ 2
2
( )
= c τ 1 + H0 + H0 (2 + q0) + · · ·
2 6

Inverting this relation, we obtain

D1 H0  D1 2 H02  D 3
1
τ= − + (1 − q0) + ···
c 2 c 6 c
c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 320
Cosmology

and the red shift is the given by

H02  D 2
z= H0 Dc1 + 2 (1 + q0) 1
c + ···

Here, D1 is the present distance (observation time) of the galaxy.

Unfortunately D1 is not directly observable, however, it is related to observable


distances.

The form of the relation exhibits the two key parameters H0 and q0 and

z = H0 Dc1

is the Hubble law.

The latter would be valid as an exact relation if the expansion would be


non-accelerated: q0 = 0 such that

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 321
Cosmology

S (t) = S (t0) (1 + (t − t0) H0) ; q0 = 0

Depending on the value of q0 we have:

v q0 > 0 the expansion is decelerated

v q0 < 0 the expansion is accelerated

From observations one presently extracts:

H0 = 74.2 ± 3.6 km/s Mpc (HST) 71 ± 4 km/s Mpc (WMAP) ; q0 = −0.60 ± 0.02

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 322
Cosmology

Cepheid Variable Stars in Spiral Galaxy NGC 3021 allow for more precise
determination of the Hubble constant
c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 323
Cosmology

Type Ia super nova reveal an accelerated expansion of our universe

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 324
Cosmology

From theory one expects that the


maximum Luminosity in the light
burst of a SN Ia has a “uni-
versal” value (standard candle).
The figure shows raw data ver-
sus corrected results, exhibiting
an absolute universal shape of
the light-curve

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 325
Cosmology

The result is more than remarkable as it implies that it is established now


that we are living in a universe with accelerated expansion. In fact, the neg-
ative value for q0 is a mind-boggling result! Observations of distant type Ia
supernovas R≫ indeed indicate that q0 is negative. An accelerating expan-
sion of the universe is in contrast to “naive” expectations: namely, that on
the cosmological scale the gravitational attraction of matter determines the
scene. An accelerating universe requires the attraction of normal matter
(including dark matter) to be more than counteracted by negative pressure
dark energy. The latter could be in the form of either quintessence or a pos-
itive cosmological constant. What is the meaning of that within the context
of cosmological models (GRT) will be discussed below.

For q0 = 0 the age of the universe is

l H0−1 = (13.7 ± 0.2) × 109 years [Hubble age]

the time back a Big Bang must have been taking place.

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 326
Cosmology

The corresponding horizon of the universe, where the escape velocity reaches the
speed of light c is

l Dmax = c
H0 ' 3.34 Gpc

about 10.9 × 109 light years.

Remember that cosmology takes place at distances D ≥ 1 Gpc, which means that
only a relatively small fraction of the universe is accessible to direct observation.
In order to see more we have to wait longer.

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 327
Cosmology

Hubble Flow: more precisely


Cosmological Redshift ⇔ The Hubble Flow cz = H0 d ⇔ due to expansion of the
Universe. To measure H0 require
l Distance
l Redshift
Must correct for local motions / contaminations
1 + z = (1 + zobs)(1 + 30/c + 3G /c)
30 = radial velocity of observer
- Measured from CMB Dipole ∼ 220 km/s
3G = radial velocity of galaxy
- contributions include Virgocentric infall,
- Great attractor etc
Decomposition of velocity field
(Mould et al. 2000, Tonry et al. 2000)
[Link to Lect.: 11,13]

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 328
Cosmology

Observable Distances
Ê Parallax Distance (for nearby objects)

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 329
Cosmology

v Observe a star six months apart,(opposite sides of Sun)

v Nearby stars will shift against background star field

v Measure that shift. Define parallax angle as half this shift

Nearest star - Proxima Centauri is at 4.3 light years =1.3 pc à parallax 0.8”
Smallest parallax angles currently measurable 0.001” à 1000 parsecs
à parallax is a distance measure for the local solar neighborhood.

Define a parsec (pc) which is simply 1 pc = 206265 AU =3.26ly.


A parsec is the distance to a star which has a parallax angle of 1”

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 330
Cosmology

First stars distances:

v α Centauri Thomas Henderson 1832 [1.35 pc]


v 61 Cygni Friedrich Wilhelm Bessel 1838 [3.48 pc]
r parallax angle the smaller the more distant the star
r atmospheric perturbations limit ground based measurements to 0.03”∼30 pc
r Hipparcos satellite: (launched by ESA in 1989 terminated 1993)
measured precision parallaxes to an accuracy of about 0.001-arcsec.

l Hipparcos measured parallaxes for about 100,000 stars


l Got 10% accuracy distances out to about 100 pc
l Good distances for bright stars out to 1000 pc.

See also: R≫, R≫

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 331
Cosmology

Parallax determination in curved space: an academic exercise

Note 1000 pc  1 Gpc curvature negligible!



S The trajectory of a light ray in the co-moving CS of

the source S is
~n

~x0 = ~n σ ; ~x0 = (r0, θ0, ϕ0)

where ~n is the direction of the ray leaving the source


at ~x0 = 0 and σ is the parameter of the corresponding
world line.
ϑ
~n1 ~nb In the co-moving system of the observer ~x = (r, θ, ϕ)
~n1
the source is located at ~x1.
b O

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 332
Cosmology

The coordinate transformation between the two systems reads:


 
√ ~
x 0
· ~
x
  q  
1
~x = ~x + ~x1 
0
,
 2


1 − k r 02 − 1 − 1 − k r

1 2 

 r1 

such that the light ray in the observer system looks like
 
√ ~
n · ~
x σ
  q  
1
~x = ~n σ + ~x1  σ
 2

1 − k 2 − 1 − 1 − k r

1 2 

 r1  

~x1
Let ~n1 = be the unit vector in direction of the source.
r1 In the flat space limit
~n = −~n1 and
 √ q 
~x = ~n σ − ~n r1 1 − k σ2 + σ − σ 1 − k r12 ,

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 333
Cosmology

where the 1st and the 3rd term cancel. Hence, the light signal hits the observer O
for σ = r1.

In weakly curved space, which is reality in our context, we are interested in signals
which pass close to O, an d which we may parametrize by the direction

~n = cos ε ~n1 + sin ε ~n⊥

with ε  1 ; ~n⊥ · ~n1 = 0 ; ~n2⊥ = 1 .

In linear approximation

~n = −~n1 + ε ~n⊥

such that
 √ q 
~x(σ, ε) ' ~n1 r1 1 − k σ2 − σ 1 − k r12 + ε σ ~n⊥

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 334
Cosmology

where σ ' r1.

Since the proper length in the RW-metric space is ` = S (t0) |~x |, for the
impact parameter b we obtain
b ' S (t0) |~x(r1, ε)| = S (t0) r1 ε

The tangent to the trajectory at σ = r1 is given by

d~x(σ, ε) ~n1

' ε ~n⊥ − q ,
dσ σ=r1

1 − k r12

and therefore the direction of observation

ε)

d~
x (σ,
q q
~nb ' − 1 − k r12 ' ~n1 − 1 − k r12 ε ~n⊥ .


dσ σ=r 1

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 335
Cosmology

With this we can calculate the angle of observation ϑ


q q
tan ϑ = 1 − k r12 ε ' 1 − k r12 S (t0b) r1 ' ϑ (ϑ  1) .

In Euclidean space with d1 as Euclidean distance of the source we would have


ϑ ' db or d1 ' ϑb .
1

Correspondingly we therefore define the parallax distance as

b S (t ) r
dP ≡ ' q 0 1 ; (ϑ, b → 0)
ϑ
1−k r12

Ë Angular Diameter Distance (distance of apparent size)

In case one has knowledge about the true size of an object, the diameter D, e.g.
from model calculations, one can determine the distance from the apparent size,
the angle ϑ:

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 336
Cosmology

Angular diameters of a) Sun b) Cartwheel Galaxy

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 337
Cosmology

D Coordinates of the source (t1, r1) in the equatorial


plane θ = π/2. The observer O is at (t0, r0 = 0).
Since dr1 = 0 ; dt = 0 the proper length

d`2 = S 2(t) r12 dϕ2


ϑ
and hence

D = S (t1) r1 · ϑ , (ϑ  1) .
O

In Euclidean space we would have d1 = D/ϑ correspondingly, one defines

D
DA ≡ ϑ ' S (t1) r1

as the distance of apparent size.

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 338
Cosmology

Ì Distance of proper motion


If one knows the true transversal velocity 3⊥, e.g., for binary stars, one can
determine the distance from the observed angular velocity of the observation
angle:
v⊥(t1 ) ∆t1 = ∆D1 First we have

S (t1)
∆D1 = 3⊥(t1) ∆t1 = 3⊥(t1) ∆t0 .
S (t0)
∆ϑ

The observed angular change per unit time is ω =


∆ϑ
ϑ̇ = ∆t . In Euclidean space we would have d1 =
0
O
3⊥ ∆t 0
∆ϑ , and correspondingly
we define the distance of proper motion as

3⊥
DM ≡ ' S (t0) r1
ϑ̇

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 339
Cosmology

The most important of the observable distances for very distant objects is the
following:

Í Luminosity Distance

This method requires the knowledge of the luminous power L1 of the source,
usually obtained form model calculations or known known physical laws or
regularities.

The power L1 of a source at (t0, r1) spreads over a surface of a sphere

4π r22 S 2(t0) .

Light which gets emitted during a time interval δt1 is received by the observer at

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 340
Cosmology

(t0, r0 = 0) during ∆t0 where

S (t0)
∆t0 = ∆t1 .
S (t1)

For the energy E = h · ν a frequency change

S (t1)
ν0 = ν1
S (t0)

is resulting, such that in total the measured Luminosity `0 is given by

1 S 2(t1)
`0 = L1 2 S 4 (t )
.
4π r1 0

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 341
Cosmology

In Euclidean space we would have `0 = L1 4π1d2 as d1 = r1.


1

Correspondingly one defines the Luminosity Distance

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 342
Cosmology

r
L1 S 2 (t0 )
DL ≡ 4π`02
= r1 S (t1) = r1 S (t0) · (1 + z)

The different observable distances satisfy the following relations:

S 2 (t1 )
DA
dL = S 2 (t0 )
= (1 + z)−2
S (t1 )
DM
dL = S (t0 ) = (1 + z)−1

We stress once more that the determination of the distances DL, DA and D M
require the knowledge of the object properties L1, D and 3⊥ (models, systematics
etc.)

One important result of the above considerations is that the proper distance D1 up
to terms of 3rd orders agrees with D M :

D1 ' D M

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 343
Cosmology

which is important for the interpretation of the red shift formula (Hubble law).

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 344
Cosmology

Comparison of Distance Measures

Proper Motion Distance:


3⊥
DM = = S (t0) r1
ϑ̇
Angular Diameter Distance:
DA = D
ϑ = S (t1) r1

Luminosity Distance:
L 1/2
 
DL = 4πF = S (t0) r1 (1 + z)

DL = (1 + z) D M = (1 + z)2 DA

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 345
Cosmology

v The Luminosity Distance (DL) shows why distant galaxies are so hard to see -
a very young and distant galaxy at red shift 15 would appear to be about 560
billion light years from us

v Even though the Angular Diameter Distance (DA) suggests that it was actually
about 2.2 billion light years from us when it emitted the light that we now see.

v The Hubble Distance (DLT) tells us that the light from this galaxy has traveled
for 13.6 billion years between the time that the light was emitted and today.

v The Comoving Distance (DcM ) tells us this same galaxy if seen today, would
be about 35 billion light years from us.

More on distance measurements R≫ R≫

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 346
Cosmology

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 347
Cosmology

History of the determination of the Hubble constant. Illustrates the problem with
cosmic distance measurements
c 2009, F. Jegerlehner
≪x Lect. 5 x≫ Credit: J.P. Hucha R≫ 348
Cosmology

Number Counts
Counting the herd. Olber’s Paradox revised. Melodie: ”Weißt du, wieviel Sternlein
stehen an dem blauen Himmelszelt” (see this: R≫)

Of particular interest:

l LUMINOSITY EVOLUTION (Galaxies in the past were brighter than today)

l DENSITY EVOLUTION (Galaxies in the past were more numerous than today)

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 349
Cosmology

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 350
Cosmology

In the co-moving CS the spatial volume element reads

1
dV 0 = √ r2 sin2 θ dr dθ dϕ
1 − k r2

and the surface element is given by

r2 dr
dVS0 = 4π √
1 − k r2

According to the cosmological principle we assume the density of the objects


(galaxies, clusters of galaxies) to be homogeneous, isotropic and time
independent (local rest system). For the presently observable universe this may
be a reasonable assumption. What we observe coming up at the horizon are
objects like the ones we know from “nearby”. In fact not quite: systematic
deviations → evolution seen!

The number of sources in the volume element dV 0 with luminosity in the interval

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 351
Cosmology

L, L + dL is

dN 0 = n0(L) dL dV 0 .

The element of the proper volume is

dV = S 3(t) dV 0 ,

such that

n(t, L) = n0(L) S −3(t)

hence

 S (t ) 3
n(t, L) = 0
S (t) n(t0, L)

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 352
Cosmology

Herewith, the number of sources at fixed cosmic time (dt = 0) in the interval
L1, L1 + dL1 in a spherical shell r1, r2 + dr1 reads

1
dN = 4 π n(t1, L1) S 3(t1) q r12 dr1 dL1 .
1 − k r12

The coordinate r1 may be replaced with the light time t1

Zt0 Zr1
c dt dr cdt1 dr1
r1 = r(t1) : = √ rsp. =−q .
S (t) 1−kr 2 S (t1 )
t1 r0 =0 1 − k r2 1

This yields the change in the number of observed sources in the time interval
t1, t1 + dt1 as
!3
S (t0)
dN = 4 π n(t0, L1) S 2(t1) r2(t1) c dt1 dL1 .
S (t1)

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 353
Cosmology

It makes sense to ask for the number of sources in a sphere, the radius of which is
fixed on the one hand by a maximum red shift zmax

S (t0)
rz = r(tz) ; tz : S (tz) =
1 + zmax

on the other hand by a minimal luminosity `min

r2(t` ) L
r` = r(t` ) ; t` : = .
S (t` ) 4 π `min S (t0)
2 4

One finds,

R∞ Rt0
N(< zmax, > `min) = 4 π c dL1 dt1 r2(t1) S 2(t1) n(t1, L1)
0 max{tz ,t` (t0 )}

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 354
Cosmology

Again we may expand in t0 − t1:

S (t1) = S (t0) {1 − H0 (t0 − t1) + · · · }


 H 0

r(t1) = S −1 (t0 − t1) 1 + (t0 − t1) + · · ·
2
n(t1, L1) = n(t0, L1) {1 − β0(L1) H0 (t0 − t1) + · · · }

Note that for the RW-metric with taking n0(L1) = constant (conservation of matter in
the co-moving local rest system) we obtain
!3
τ2
( )
S (t0)
n(t1, L1) = n(t0, L1) ' n(t0, L1) 1 + 3 H0 τ + 3 H02 (4 + q0) + · · ·
S (t1 2

which yields

β0(L1) = −3

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 355
Cosmology

In a so called steady-state theory one would have β0(L1) = 0. Of course on must


try to determine β0(L1) by observation, which is not an easy task, however.

In expanded form we obtain

Z∞

N(< zmax, > `min) ' dL1 n(t0, L1) (t0 − tmax(L1))3
3
0
3
× {1 − (β0(L1) + 1) H0 (t0 − tmax(L1)) + · · · }
4

where tmax = max {tz, t` (L1)}.

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 356
Cosmology

In spite of the fact that new telescopes, in particular the Hubble space telescope,
have dramatically increased the number of observed objects, it remains very
difficult to determine the parameters β0(L), specific for number counts, and H0 or
even q0 just from number counts. But any kind of information is needed. In case of
radio sources one does not measure the total luminosity but the spectral
distribution. The above formulas have to be modified to include the appropriate
frequency shifts as well (see S. Weinberg I).

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 357
Cosmology

v Study Galaxy source counts to

à investigate the contribution of different galaxy populations to the Universe


à compare the evolution of galaxies today with those in the past
à constrain Geometry of the Universe

v Source counts depend on

à Cosmology
à Luminosity Function
à K-Correction
à Evolution

v Source counts are a function of observing wavelength

à Different wavelengths are dominated by different classes of sources


à To understand the star formation and evolutionary history à multiwavelength
analysis

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 358
Cosmology

v Present and next generation surveys


à Larger areas to greater depths
à increase statistical samples by orders of magnitudes!

Left: Differential UBVIJHK galaxy counts as a function of AB magnitudes. Note

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 359
Cosmology

the decrease of the logarithmic slope d log N/dm at faint magnitudes. The
flattening is more pronounced at the shortest wavelengths. Right: Extra-galactic
background light per magnitude bin, (i = 10-0.4(mAB+48.6) N(m), as a function of
U (filled circles), B (open circles), V (filled pentagons), I (open squares), J (filled
triangles), H (open triangles), and K (filled squares) magnitudes. For clarity, the
BVIJHK measurements have been multiplied by a factor of 2, 6, 15, 50, 150, and
600, respectively.

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 360
Cosmology

More here:R≫ R≫

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 361
Cosmology

The Hubble Space Telescope (HST) has been upgraded recently and continues to
look deeper into space and time.
c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 362
Cosmology

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 363
Cosmology

An “empty spot” region ...

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 364
Cosmology

... looked at closer (using new IR cameras)

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 365
Cosmology

References

v S. Weinberg, Gravitation and Cosmology, Wiley 1972

v S. Weinberg, Cosmology, Oxford 2008

v M. Roos, Introduction to Cosmology, Wiley 2003

v D. Harland, The Big Bang, Springer 2003

v J. Rich, Fundamentals of Cosmology, Springer 2001

v L. Bergström, A. Goodbar, Cosmology and Particle Physics, Springer 2004

v E. Harrison, Cosmology, Camb. Univ. Press 2000

[back]

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 366
Cosmology

Appendix: spaces of constant curvature


Theorem 3. Constant Gauss curvature in a point P

Rµν,ρσξµξρηνησ
K(P) =   = constant
gµσgνρ − gµρgνσ ξµξρηνησ

./
 
Rµν,ρσ = K gµσgνρ − gµρgνσ .

Proof: 1) We define the auxiliary tensor


 
T µν,ρσ  Rµν,ρσ − K(x) gµσgνρ − gµρgνσ ,

where K = K(x) may depend on the point x. By construction T µν,ρσ has the

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 367
Cosmology

index symmetries of Rµν,ρσ, which are

Rµν,ρσ = −Rµν,σρ ; Rµν,ρσ = −Rνµ,ρσ


X
Rµν,ρσ = Rρσ,µν ; Rµ[ν,ρσ] = 0 ; [. . .] : .
cycl.

In compliance with the requirements

T µν,ρσξµξρηνησ ≡ 0

holds for arbitrary vectors in the tangent space at P. The curvature K(x) at P is
direction independent (isotropy). We may take special vectors, with two
independent components:

ξµ = a δµα + b δµβ
ηµ = c δµγ + d δµδ

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 368
Cosmology

and consider the term proportional to abcd:

T αγ,βδ + T βγ,αδ + T αδ,βγ + T βδ,αγ = 0 ,

where in the 2nd and 4th term by symmetry we may interchange the index pairs:
T βγ,αδ = T αδ,βγ etc. Thus

T αγ,βδ + T αδ,βγ = 0 .

By a different choice of the vectors similarly we obtain

T αγ,δβ + T αβ,δγ = 0 .

Now we consider the “Jacobi-Identity” (it holds for R... and hence for T ...):

T α[γ,βδ] ≡ T αγ,βδ + T αδ,γβ + T αβ,δγ = 0 .

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 369
Cosmology

Using the two relations derived above together with the antisymmetry of R... and
hence T ... in the second index pair we find: the 2nd term is T αδ,γβ = −T αδ,βγ = T αγ,βδ
and the 3rd T αβ,δγ = −T αγ,δβ = T αγ,βδ and therefore

T α[γ,βδ] ≡ 3 T αγ,βδ = 0

for any combination of the indices, i.e. T ... ≡ 0 q.e.d. q

Corollary 1. An isotropic space has a curvature tensor

 
Rµν,ρσ = K(x) gµσgνρ − gµρgνσ

Proof: 2) For the isotropic space the Ricci-tensor reads:

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 370
Cosmology

Rµν = Rλµ,νλ = gλω Rωµ,νλ


ωλ
 
= K(x) · g gωλgµν − gωνgµλ
= K(x) · (n − 1) gµν ,

and together with the scalar curvature

R = gµν Rµν = K(x) · n (n − 1) ,

the Einstein-tensor obtained is

1 (n − 2)(n − 1)
Gµν = Rµν − R gµν = −K(x) gµν .
2 2
µ
The Bianchi-Identity: Dµ G ν = 0 infers

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 371
Cosmology

K(x) is a scalar field


 µ
and hence Dµ K(x) = ∂µ K(x), also Dµ δ ν ≡ 0, such that

1
Dµ Gµν = − (n − 2)(n − 1)∂ν K(x) ≡ 0
2

i.e. K(x) = K = constant q.e.d. q

Corollary 2. For n ≥ 3 an isotropic space is also homogeneous

K(x) = K = constant (n ≥ 3)!

Theorem 4. (uniqueness) Two Riemann spaces M n and M n with the same con-
stant curvature are locally (in neighborhoods of given points P and P) isometric.

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 372
Cosmology

Proof: By assertion there exists a mapping

∂xµ
Λµν(x)  ν
∂ x̄

such that

gµν Λµρ Λνσ = ḡρσ . (∗)

The latter “mapping” in fact is a system of differential equations which is to be


solved. By linear transformations one can achieve that in P and P the initial
conditions

gµν(0) = ḡµν(0) =gµν and Λµν(0) = δµν

hold. The integrability condition for (*) we obtain by differentiation of it. Denoting
∂λ ≡ ∂∂x̄λ and utilizing the transformation laws for the Christoffel symbols,

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 373
Cosmology

we obtain

∂λΛµν = Γ̄ανλ Λµα − Γµβγ Λβν Λγλ (∗∗)

For each solution of (**)

ḡρσ − gµν Λµρ Λνσ = Cρσ = constant

holds, where by the choice of the initial conditions

Cρσ ≡ 0 .

Together with (**) we have

∂2 x µ
∂λ Λµν = λ ν = ∂ν Λµλ
∂ x̄ ∂ x̄
c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 374
Cosmology

where the symmetry is the integrability condition for

∂xµ
Λµν = ν.
∂ x̄

Now, the integrability condition for (**) is precisely

R̄αβ,γδ = Rµν,ρσ Λµα Λνβ Λργ Λσδ .

However,
 
R̄αβ,γδ = K ḡαδḡβγ − ḡαγ ḡβδ

and
 
Rµν,ρσ = K gµσgνρ − gµρgνσ

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 375
Cosmology

are the preconditions of the theorem, such that because of (*) the integrability
conditions indeed are satisfied.

q.e.d. q

Corollary 3. In a Riemannian space of constant curvature to two given points P


and P0 there exists an isometric mapping which maps a neighborhood of P into a
neighborhood of P0.

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 376
Cosmology

Explicit construction

Special forms of the metric. Starting point is

M n+1: n+1-dimensional Euclidean or pseudo-Euclidean space


ds 2
= gAB dxAdxB ; A, B = 1, · · · , n + 1

x n+1
z ; gAB= diag (1, · · · , 1, k) ; k = ±1, 0
 2
ds2 = dxi + k dz2 = hik dxidxk ; i, k = 1, · · · , n

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 377
Cosmology

Hypersurface: xi xi = ρ2 − k z2 defines an isometric embedding

k = 1 , ρ2 > 0


 sphere
xi xi = ρ2 − k z2 ;  k = −1 , ρ2 < 0

two sheet hyperboloid

 k = −1 , ρ2 > 0


one sheet hyperboloid

Take k , 0: k z2 = ρ2 − x2  2 k zdz = −2 xidxi thus

 2  2
i i
xidx 1 xidx
dz =
2
= ; k 2
=1
k 2 z2 k ρ2 − x 2

and

 2
i
xidx
i 2 i 2
   
ds = dx
2
+ k (dz) = dx 2
+ = hik dxidxk
ρ2 − x2

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 378
Cosmology

xi xk xi xk
hik = δik + 2 ; h =δ + 2
ik ik
ρ −x 2 ρ

Check:

xi xl x i xl xi xl x2
h hkl = δ l − 2 + 2
ik i
− = δi
l 3
ρ ρ −x 2 ρ (ρ − x )
2 2 2

The derivative of the metric is


( )
1 xi xk xl
∂l hik = xi δkl + xk δil + 2 2
ρ2 − x2 ρ − x2
1
= (xihkl + xk hil)
ρ −x
2 2

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 379
Cosmology

and the Christoffel symbol reads

1
Γi,kl = (∂lhik + ∂k hil − ∂ihkl)
2
1
= xihkl
ρ −x
2 2

1
Γ jkl = h jiΓi,kl = 2 x jhkl
ρ

1  j 
∂r Γklj = δr hkl + x ∂r hkl
j
ρ2

1   j 
= ρ − x δr hkl + x xk hlr + x xlhkr
2 2 j j
ρ (ρ − x )
2 2 2

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 380
Cosmology

Ri k,lm = ∂m Γi kl − ∂l Γi km + Γi λm Γλkl − Γi λl Γλkm


1  i 
= δ mhkl − δ lhkm ,
i
ρ 2

1
Rik,lm = (himhkl − hilhkm)
ρ2

As a result
 space of constant curvature
K = ρ12

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 381
Cosmology

z k= 1 z k = −1
ρ2 < 0
ρ2 > 0 ρ2 > 0
|ρ| i|ρ|

|x| |ρ| |x|

z 2 + x2 = ρ2 −z 2 + x2 = ρ2

Corollary 4. Riemann spaces of constant curvature exist for arbitrary K ∈ R. Ac-


cording to the previous theorem these are locally unique.

Result: Isotropic spaces are homogeneous and of constant curvature and


coordinates may chosen locally such that the metric is

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 382
Cosmology

 2 ( xidxi)2
ds2 = dxi + ρ2 −x2

Symmetries of spaces of constant curvature

The metric derived above per construction possesses an invariance group which is
induced by the orthogonal or pseudo-orthogonal group on M n+1. The line element

ds =gAB dxAdxB
2

as well as the embedding condition

x A x A = ρ2

are invariant under

x0A = ΛAB xB ; ΛAB real ,

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 383
Cosmology

where
◦ ◦
gAC ΛAB ΛCD =gBD .

If we write
!
R a ◦ ◦
Λ= ; Λ g Λ =g ,T
b c

the requirement of the invariance of the metric tensor can be split into the
conditions:
 
RR T
+ k aiak = δik ,
ik

Rik bTk + k c ai = 0 ,
bibi + k c2 = k .

On the induced space of constant curvature, we can distinguish two types of

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 384
Cosmology

transformations:
Ê Rotations:
ai = 0  R RT = 1  bi = 0 ; c = 1
Evidently,
 2
i
 2 xidx
ds = dx + 2
2 i
and k z2 + x2 = ρ2
ρ −x 2

both are invariant under x0i = Ri k xk . These rotations leave the origin xi = 0
invariant.
Ë Quasi-translations:
ai , 0  bi , 0 ; c = 1 − bibi/k
p

We are looking for a matrix

Rik  δik − rik

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 385
Cosmology

which for ai = 0 goes to Rik = δik . This requires rik to be of the form

rik = λ k aiak .
 
Then R R T
+ k aiak = δik ; k2 = 1 implies
ik

−2λk aiak + λ2a2aiak + k aiak = 0

and hence

λ2 a2 − 2 λ k + k = 0 ./ λ2 k a2 − 2 λ + 1 = 0 .

Thus

1  √ 
λ± = 2 1 ± 1 − k a ; a2  aiai
2
ka
c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 386
Cosmology

where only λ− is regular at a2 = 0 and hence

 √ 
λ ≡ λ− = 1
k a2
1− 1− k a2 .

Furthermore, from Ri k bTk + k c ai = 0 we obtain

bTi − λ k aiak bTk + k c ai = 0 (R)

with solution
bTi = −k ai .

Here, we have used c = 1 − bib /k = 1 − k a2 and (R) yields
p
i

−1 − λ k a2 + c = 0  λ = 1−c
k a2
.

Rik = δik − λ k aiak


Altogether: bTi = −k
√ ai
c = 1 − k a2 ; λ = 1−c
k a2

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 387
Cosmology


where ai is an arbitrary n-dimensional vector with the constraint that c = 1 − k a2
is real! Therefore, for

x0i = Ri k xk + ai z where k z2 = ρ2 − x2

we have
q !
x0i = xi + ai |ρ2 − x2| − λ k x · a

This transformation in particular transforms the origin xi = 0 into the vector


ai
x0i(0) = ai |ρ|, that’s why we call it quasi-translation. If we set |ρ| = āi we finally
obtain: √ 
x0i = xi + āi 1 − K x2 − λ K x · ā

where K = k/ρ is the Gauss curvature and ā is such that c̄ = c =
2 i
1 − K ā2 is
real and λ̄ = λ = K1−c
ā2
. That’s it!

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 388
Cosmology

Previous ≪x , next x≫ lecture.

c 2009, F. Jegerlehner
≪x Lect. 5 x≫ 389

You might also like