You are on page 1of 23

A comparison of the different cosmological models of the

universe
Aryan Gadhave
January 2023

1
Contents
1 Abstract 3

2 Introduction 4
2.1 What is Cosmology? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.2 A brief history of Cosmology before Newton . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

3 Newtonian Cosmology 6
3.1 The cosmological principle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
3.2 Modelling the expansion of the universe . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
3.3 Value of the k term . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
3.4 The fluid and acceleration equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
3.5 A comparison of matter and radiation dominated universes . . . . . . . . . . . . . . . . . . . . . 8

4 General Relativity 9
4.1 Problems with Newtonian Gravity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
4.2 The equivalence principle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
4.3 Gravity and Curvature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
4.4 Tensors in General Relativity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
4.5 The Metric Tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
4.6 The geodesic equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
4.7 The Riemann Curvature Tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
4.8 The Ricci Tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
4.9 The Energy Momentum Tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
4.10 Curved Newtonian gravity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
4.11 The Einstein Field Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

5 The FLRW metric 14


5.1 The Universe as a Perfect Fluid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
5.2 The Cosmic Rest Frame . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
5.3 The Friedmann equations with dark energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
5.4 A comparison of the different solutions to the Friedmann equations . . . . . . . . . . . . . . . . . 16
5.5 Limitations of the solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

6 Conclusion and Remarks 17

7 Appendix 18
7.1 Vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
7.2 Co-vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
7.3 The Tensor Product and the general definition of a tensor . . . . . . . . . . . . . . . . . . . . . . 19
7.4 Derivation of the geodesic equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
7.5 The Covariant derivative . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
7.6 Riemann Tensor components . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
7.7 Calculating the Christoffel symbols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
7.8 Proof of the divergence-free property of the Einstein tensor . . . . . . . . . . . . . . . . . . . . . 21
7.9 The metric tensor for flat space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
7.10 The metric tensor for a closed/spherical space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
7.11 The metric for hyperbolic/open space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
7.12 Unifying the 3 metrics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

References 22

2
1 Abstract
This paper will examine how cosmology has developed over the course of human history- from
the geocentric model of the Ancient Greeks to Kepler’s revised heliocentric model, the Newto-
nian model for Cosmology and finally, the incorporation of General Relativity. In particular,
the latter two will be focused more heavily on. Although the use of mathematics will be kept
accessible, certain things such as a basic understanding of linear algebra and calculus will be
assumed, particularly for Sections 4 and 5. In the final section, there will be a discussion of
large-scale geometries of the universe- flat, spherical and hyperbolic, as well as solving the
Friedmann equations with dark energy.

3
2 Introduction
2.1 What is Cosmology?
There are few fields of study as inclusive as cosmology, which itself could be defined as the study
of the physical properties of the universe. However isn’t this just physics in general? Avoiding
too broad a definition leads us to the following; cosmology is the study of the observable
universe’s origin, its large-scale structures and dynamics, and the ultimate fate of the universe.

2.2 A brief history of Cosmology before Newton


One of the very first cosmological models for the solar system was developed by the ancient
Greeks and is known as the geocentric model- where the geo refers to Earth, i.e. assuming Earth
is at the centre of the Solar System. However, there are several reasons for this being a very
primitive model. Firstly, if you assume the geocentric model is correct, then Mars appears to
reverse its orbital direction from time to time. To correct this Ptolemy, the Greek philosopher,
introduced epicycles which are small circular motions by planets on top of their circular motion
with Earth (Figure 1). As further astronomical data was gathered, the models were revised by
adding more epicycles, until they became so complex that they were completely unusable.

Figure 1: The path of Mars according to Ptolemy’s epicycle model- highlighted in red [11]

Although the geocentric model remained the dominant viewpoint for over a 1500 years, its
antithesis, the heliocentric model- meaning sun at the centre, was also proposed around 270
BC by Aristarchus however it failed to gain traction until the time of Copernicus for a number
of reasons. Firstly, it implies that Earth is in a circular orbit around the Sun meaning its
in constant motion, but at the time of ancient Greece- motion that wasn’t ’felt’ was deemed
impossible. Furthermore, according to the model there must be parallax between Earth and
nearby stars other than the Earth (which is in fact true), but due to the vast astronomical
distances the angle of parallax was so minute that it wasn’t experimentally measured until
1838.

Copernicus’s (1473-1543) work vastly advanced the heliocentric model through the data he
collected however it was still ultimately flawed and it wasn’t until 1609, when Kepler published
his empirical laws of planetary motion that, a theory which accurately depicted physical reality

4
and made robust predictions was developed. Kepler’s 1st law stated that the planets orbited
the sun in ellipses by with an eccentricity ϵ, which can be thought of as a measure of how much
an ellipse deviates from a perfect circle. Kepler’s 2nd law states that planets sweep out an
equal area over equal time intervals during their orbits. Kepler’s 3rd law is however the most

Figure 2: Demonstration of Kepler’s 2nd law where A1 = A2 , and ai are the semi major axes and Fi are the
foci [6]

notable and useful for predictions; it states that T 2 ∝ a3 where T is the orbital time period
and a is the length of the semi-major axis.

Kepler’s laws are what lead Newton to the law of gravitation (3.2), which is remarkable for
its predictive power given its simplicity.

5
3 Newtonian Cosmology
3.1 The cosmological principle
This states that the spatial distribution of matter in the universe is isotropic and homogeneous,
meaning it is uniform throughout space and in all directions (rotationally symmetric). Exper-
imental data seems to support this viewpoint and it forms the foundation of cosmology.
This principle also motivates the definition of the scale factor; a(t) = a0 R(t)
R0
where a0 and R0
are the radius and scale factor at an arbitrary time, t0 , usually taken to be the age of the
universe. This is particularly useful as when forming a Newtonian theory of cosmology, you
can simply take a point in the universe, O and create a sphere of radius R0 and then by the
cosmological principle, the evolution of the sphere is a proxy for the evolution of the universe.

3.2 Modelling the expansion of the universe

Figure 3: The experimental data used by Hubble to derive his law [1]

As per the cosmological principle, our expanding sphere must have a uniform density, ρ and

thus its mass M is as follows: M = 4πR(t) 3
. Also, the expansion is homoglous, meaning the
mass M is constant. Another key characteristic is the velocities of objects within the sphere
due to the expansion are radial and given by Hubble’s law (Figure 3), v(t) = H(t)R(t) or
equivalently Ṙ
R
= ȧa = H using the definition of the scale factor.
2
By the law of the conservation of energy, the total energy E = mv2 − GM m
R(t)
(sum of the kinetic
energy and gravitational potential energy) must remain constant as the sphere expands radially.
2Ea2
Substituting the expressions for v and M and rearranging for H 2 yields H 2 = 8πGρ3
+ ma(t)20R2 .
0
2E
Given that mR 2 is a constant, when working in the context of general relativity, it is convenient
0
to replace it by −Kc2 where c is the speed of light in a vacuum. This along with the relation
kc2
Ka20 = k where k can only equal 0, 1 or -1 yields the Friedmann equation H 2 = 8πGρ 3
− a(t)2 (1),

describing the expansion of the universe.

6
3.3 Value of the k term
According to the Friedmann equation, the value of the k/curvature term will determine the
ultimate fate of the expansion of the universe. The value of k = −1 corresponds to negative
curvature and the value of ȧ is always positive meaning the universe will continue expanding
forever. Also in terms of the large scale geometry, it corresponds to a hyperbolic, open universe
(Figure 4). On the other hand, if k = 1 then eventually the value of ȧ will be negative and
the universe will start to contract. This corresponds to a closed, spherical universe ultimately
ending in a Big Crunch. Finally, if k = 0 the universe will continue to expand however the rate
of expansion will tend to 0 with time since the mass density, ρ will tend to 0. Also k = 0 means
that there is no overall curvature and the large scale structure of the universe is flat, which is
the most likely scenario according to observations.

Figure 4: Large scale geometries of the universe [12]

3.4 The fluid and acceleration equations


Looking back at the Friedmann equation, it is impossible to determine the scale factor for our
universe through the sole use of (1) since the mass density, ρ(t) and the scale factor are both
undetermined functions of time. To explicitly find the scale factor and mass density, you need
another equation linking the two parameters.

The first law of thermodynamics states that the change in heat flow dQ is the sum of the
change in internal energy dU and change in work done dW . It can be rewritten in a more
convenient form as follows: dQ = dU + pdV . Here the total internal energy in our sphere of
4πa(t)3 R03 ρ(t)c2
mass M can be expressed as, U (t) = M c2 = 3
. Differentiating w.r.t time yields:
dU
dt
= 4πR03 c2 a(t)2 ρ(t)ȧ + 43 πR0a 3c2 a(t)3 ρ̇. Meanwhile you also have dV
dt
= 4πR03 a(t)2 ȧ. Fi-
nally, as per the cosmological principle, our sphere is perfectly homogeneous and so the net
heat flow dQ = 0. Substituting our expressions yields the fluid equation: ρ̇ + 3 aȧ ( cp2 + ρ) = 0 (2).

Going back to (1), rearranging and differentiating w.r.t. time gives: äa = 4πG 3
(2ρ + ρ̇ aȧ ).
Finally, substituting for ρ̇ gives the acceleration equation: aä = − 4πG
3
(ρ + 3p
c2
) (3).

There are quite a few notable things about (3). Firstly, you see that an increase in pressure
p contributes to an overall deceleration of the expansion of the universe due to the gravitational
force. The contribution from the mass density, ρ is more obvious since the addition of matter/
mass slows down the outward expansion due to the attractive force of gravity. Also, the
curvature term k does not appear in (3), meaning that irrespective of the large scale geometry
of the universe, the universe cannot keep expanding at an accelerating rate. However, this is
in a direct contradiction with observations, meaning that (3) is incomplete. When Einstein

7
Λ
encountered this, he added the infamous cosmological constant by adding a 3
term, which
accounts for dark energy, driving the universe’s accelerated expansion.

3.5 A comparison of matter and radiation dominated universes


In Cosmology, there is an underlying assumption that there is a unique pressure associated
with each mass density, or p = f (ρ) which is known as the equation of state. This can be
applied to the two key constituents of the universe: dust and radiation. Firstly, dust refers to
non-relativistic matter, such as galaxies, and in general matter in the form of atoms. You find
that for dust the equation of state reduces to p = 0. Secondly, radiation refers to relativistic
2
particles such as photons of light which have a corresponding equation of state, p = ρc3 , known
as radiation pressure (the same pressure generated from nuclear fusion in stars). Although this
model is inherently flawed due to the existence of dark matter and energy, it can be used to
model the evolution of a simplistic universe where dark matter and energy do not exist.

Firstly, considering the Friedmann and fluid equations for a flat universe composed of dust,
2
you have: ȧ2 = 8πGρa
3
(1). On the other hand (2) yields ρ̇ + 3 ȧa ρ = 0 which can then be solved
ρ0 t20
to give ρ = aρ03 . Substituting into (1) yields the solution, a(t) = ( tt0 )2/3 and ρ(t) = t2
. This
describes a universe that expands forever however at a decelerating rate. [8]

Figure 5: Graph of ρm = ρdust and ρrad against time [9]

On the other hand, if the universe is dominated by radiation instead (the first 47000 years
after the Big Bang) you find that the solution to the Friedmann and fluid equations is as
ρ t2
follows: a(t) = ( tt0 )1/2 and ρ(t) = aρ04 = t02 0 . This means that the universe expands slower if
it is composed of radiation due to the radiation pressure term in (3). Furthermore, ρdust ∝ a13
whereas ρrad ∝ a14 so the radiation solution isn’t stable and eventually, the universe will be
dominated by dust which itself is stable as time passes.

8
4 General Relativity
4.1 Problems with Newtonian Gravity
Perhaps most obviously, the biggest problem with Newtonian gravity is that it is a non-
relativistic theory of gravity and so is accurate only for cases with weak gravity and low veloc-
ities. Also, the interpretation of the gravitational force is that it is instantaneous which means
information is transmitted faster than the speed of light, violating a key postulate of relativity.
As per Newtonian gravity if the Sun suddenly disappeared, all the planets would be flung out
of orbit instantly, however in reality it would take around 8 minutes for it to happen on Earth
(the gravitational wave would have to travel at the speed of light).

A second yet more subtle effect is the precession of Mercury’s orbit around the Sun. In fact,
this effect occurs to all the planets in the Solar System but is only prevalent in Mercury’s orbit.
The perihelion of a planetary orbit is defined as the point on the ellipse of closest approach,
and in a precession, this point varies over time. This is again in contradiction with Newtonian
gravity which produces elliptical orbits without precession.

Figure 6: Mercury’s orbit around the Sun [15]

4.2 The equivalence principle


The principle of relativity, at its core, is about how special inertial reference frames are- which
themselves can be thought of as non-accelerating co-ordinate systems. The statement of the
principle is that, any identical physical experiments carried out in two different inertial frames
will yield the same result. Namely, when this is applied to gravity you get the Einstein equiv-
alence principle; the outcome of any local non-gravitational experiment in a freely falling lab-
oratory is independent of the velocity of the laboratory and its location in space-time. Here,
locality assumes there is no way to look outside, i.e. the laboratory is in a black-box. Also,
the fact that the outcome is independent of the location in space-time implies that if I have
two observers Alice and Bob, and I put Alice on a spaceship accelerating upwards at 9.81ms−2
and Bob on the surface of the Earth, if they are both in a local environment then it would be
impossible for them to know if they were on Earth or on the spaceship, or in another equiv-
alent system. Said another way a gravitational field and an accelerating reference frame are
physically identical.

9
Figure 7: Alice and Bob throw a tennis ball, it has an identical trajectory as suggested by the equivalence
principle

However there is one and very major flaw in this argument- the assumption that there are
no tidal forces, which occur due to variations in the gravitational field. In reality, Bob might
actually be able to tell that he is on Earth, since even close to the surface, Earth’s gravitational
field isn’t perfectly homogeneous. This is why when the equivalence principle is used to derive
any further physical laws, it must be in a system with minimal tidal forces.

4.3 Gravity and Curvature


The equivalence principle, very remarkably, also shows how gravity is related to curvature. To
see this consider Alice and Bob from earlier, but this time Alice is in a box in space in a region
of zero gravitational field and Bob is falling towards the surface of the Earth, also in a box.
Assuming no tidal forces the observers are physically equivalent, they both ’feel’ weightless.
Now imagine they both have a laser and shine it across the box horizontally, then the result
must be identical and from each of their perspectives, the beam of light travels in a straight
line across the box. As Bob falls, what does an inertial observer see? Imagine someone from
the surface of the Earth could look inside Bob’s box. Then they must see the light bend due
to the relative motion of free fall. Or equivalently, the beam of light when observed from an
inertial reference frame seems to curve towards the surface of the Earth, due to the presence of
its gravitational field.
To make this curving effect seem even more prevalent. Imagine you have three observers in flat
space-time and you set them out in an equilateral triangle and they send pulses of light to one
another. They can then measure the angles and will confirm that they add to 180°. However,
this can only hold in flat Euclidean geometries, so now imagine adding a black-hole in between
the observers. Now when they send the pulses of light, they will curve and thus lead to the
sum of the angle measurements to exceed 180°.

Figure 8: The two triangles show the paths taken by light in each scenario

4.4 Tensors in General Relativity


Tensors are mathematical objects defined by the way they transform under a change of basis
(analogous to changing coordinate systems). They play a key role in the way general relativity

10
is formulated as they allow you to explore the geometry of space-time and detect the presence of
gravity, or equivalently curvature. The precise definition of a tensor (see 5.1) is mathematically
involved whilst not being physically intuitive, therefore it can be convenient to think of them
as matrices with components obeying a special set of transformation rules.

4.5 The Metric Tensor


The metric tensor can loosely be thought of as a generalisation of the gravitational potential
from Newtonian gravity, appearing in the Einstein field equations. Its components are defined
as follows, gµν = e⃗µ · e⃗ν and can be written in a matrix, which for a Cartesian basis, reduces to
the identity matrix. Equivalently it can be thought of as a function taking two vector inputs
and producing a real number, or g(⃗v , w) ⃗ = v i wj (⃗
⃗ = ⃗v · w ei · e⃗j ) = v i wj gij with summations over i
and j. The metric plays a key role in general relativity due to the fact that knowing the metric
components allows you to calculate distances in curved geometries, where the Pythagorean
theorem no longer applies. To see this consider a vector ⃗v , and then you have ∥v∥2 = ⃗v · ⃗v =
v i v j (⃗
ei · e⃗j ) = v i v j gij . This means that for a given space, the metric tensor captures all
information about its geometric structure; many other tensors are also defined in terms of the
metric, making it a fundamental object of study in general relativity. [5]

4.6 The geodesic equation


As suggested by the equivalence principle, a free-falling object is physically equivalent to an
object in space with a zero gravitational field, due to the feeling of ’weightlessness’. Both these
objects travel on so-called geodesics, which are curves that describe the natural motion of an
object through space-time when no forces are applied. Therefore, a geodesic on a surface can be
intuitively thought of as the straightest possible path through that surface, and more formally
a path without any tangential acceleration, or where the acceleration vector is only normal to
the surface. In the earlier examples, Alice has no acceleration vector, and Bob’s acceleration
vector points directly towards the centre of Earth i.e. it acts normally, therefore both their
paths can be considered geodesics.
In general relativity, the geodesic equation is really a set of 4 differential equations (for each
of the space-time components X µ ) in terms of the Christoffel symbols Γijk which are the coef-
ficients of the acceleration/normal vector. It can then be solved to determine the trajectory of
a particle in terms of the parameter T . It is a formalisation of the equivalence principle, and
therefore can be used to derive the precession of Mercury as well as the gravitational lensing of
light (4.1).

4.7 The Riemann Curvature Tensor


The curvature on a surface is formally defined in terms of the Riemann Curvature tensor, which
uses the concept of parallel transporting vectors- transporting vectors such that they remain as
straight as possible. It can be thought of as a way to generalise Gaussian curvature to 3 dimen-
⃗ and ⃗t then form a parallelogram
sions or higher. The idea is as follows, you take 3 vectors ⃗v , w
using w ⃗
⃗ and ⃗v . The vector t is then transported along the loop formed by the parallelogram to
get another vector ⃗t˜, and the vector difference is the output of the Riemann curvature tensor.
If this vector difference is non-zero then the surface is curved, and changes in the Riemann
curvature tensor allow you to track changes in curvature; determining the large-scale geometry
of a space. The Riemann tensor is formally defined in terms of the Christoffel symbols as well
as the covariant derivative (7.5), which are the mathematical analogues of parallel transport.
ρ
In 4-D the Riemann curvature tensor Rσµν has 256 components and calculating all of these is
computationally intensive. Fortunately, there are symmetries between the components, mean-

11
ing that in 4-D there are 20 independent components. However, sometimes it is still convenient
to work in terms of the Ricci tensor, thought of as a summary of the Riemann tensor.

4.8 The Ricci Tensor


c
Motivated by the definition from earlier you have the Ricci tensor, Rac = Racb , where you
sum over the middle and top components of the Riemann tensor. Similarly to the Riemann
tensor, the Ricci tensor also has geometric meaning as it tracks volume changes along geodesic
curves in space-time. This is analogous to the effect of gravity in General Relativity as objects
are attracted due to their geodesic paths converging as a result of the curved local geometry
of space-time. Formally, you have the Ricci curvature, Ric(⃗v , ⃗v ) = v j v i Rij and if for a path
along the vector, ⃗v all the geodesics converge, Ric(⃗v , ⃗v ) > 0, and if they diverge Ric(⃗v , ⃗v ) < 0.
Equivalently the volume of an object, all of whose points travel on geodesics will decrease if
Ric(⃗v , ⃗v ) > 0 and will increase if Ric(⃗v , ⃗v ) < 0. However, the case of Ric(⃗v , ⃗v ) = 0 is more
interesting since it means on average the geodesics along ⃗v are invariant, although individual
geodesics may deviate. Likewise, the volume remains invariant although the shape of an object
may change as it moves along the geodesics. Finally the Ricci scalar, R = g µν Rµν measures
how the size of a sphere deviates in a curved geometry compared to that of a flat one. The
Ricci tensor and scalar appear on the LHS of the Einstein field equations and describe how the
geometry of a space influences the path of objects, i.e. the effect of gravity.

4.9 The Energy Momentum Tensor


General Relativity describes gravity as synonymous with the curvature of space-time, and this
curvature is a result of the presence of mass, or equivalently energy. In the context of Newtonian
gravity, the presence of mass and energy was represented by ρ the mass density. However, this
quantity isn’t Lorentz invariant, meaning it can change depending on your frame of reference.
To demonstrate this consider a unit box with mass density in its rest frame ρ. As it moves, its
sides are length contracted, so the mass density falls. For General Relativity this object can be
generalised to form the energy-momentum tensor, T µν . Its components are defined as follows,
T µν = the flux of µ momentum going through a unit box with constant ν i.e. holding the X ν
co-ordinate constant and tracking how many field lines of the µ-th component of momentum
pass through the unit box spanned by the other 3 co-ordinates.

Figure 9: The energy-momentum tensor as a matrix

For example, if time t is held constant then the components T αt track the α component
of the momentum going through a unit box of ∆x, ∆y, ∆z. This means that, the time-time
Pt E
component T tt = ∆x∆y∆z = ∆V = ϵ = ρc2 is the energy density.

Likewise, if x is held constant T αx = ∆y∆z∆t which for α ̸= t correspond to pressure and
shear stress terms, responsible for rotating and squishing objects.

12
4.10 Curved Newtonian gravity
It turns out that Newton’s law of universal gravitation can be generalised using the Ricci and
energy-momentum tensor. You have F⃗ = − GM r2
m
e⃗r = m⃗g ⇔ ∇ · ⃗g = −4πGρ. Using the fact
that ⃗g = −∇ϕ where ϕ is the gravitational potential yields ∇ · ∇ϕ = ∇2 ϕ = 4πGρ (Poisson’s
equation), where the ∇2 is the Laplacian. This can be expressed in terms of the Ricci ten-
2 i ∂ϕ
sor as follows: firstly the equation of motion using this becomes ⃗a = ⃗g ⇒ ddtx2 e⃗i = − ∂x ie
⃗i .
2
d x i ∂ϕ
Equating the components yields dt2 + ∂xi = 0 and comparing this to the geodesic equation
∂ϕ
allows you to compute the Christoffel symbols giving Γi00 = ∂x i with all other coefficients being

zero. Finally, the Riemann and Ricci tensor components are then computed using the Christof-
fel symbols and you have R00 = ∇2 ϕ with all other components equal to zero for the Ricci
tensor. Also, in the context of Newtonian gravity the energy-momentum tensor only has one
non-zero term, the energy density. Poisson’s equation then becomes R00 = 4πG c2
T00 (6) which
hints at the form of the generalised Einstein field equations as an equivalence of the curvature of
space-time/gravity to the presence of mass/energy in the form of the energy-momentum tensor.

Therefore, curved Newtonian gravity can be thought of as an intermediary theory between


Newtonian gravity and General Relativity, since it incorporates the use of space-time as being
equivalent to gravity as well as using the geodesic equation to solve for the path of objects.
However, it is still a non-relativistic theory, meaning it uses the classical Galilean transformation
laws i.e. assuming time is absolute/universal, the constancy of the speed of light isn’t obeyed.

4.11 The Einstein Field Equations


The main problem with (6) is that it is a non-relativistic equation, however, it does suggest
a generalisation, namely that Rµν = κTµν with κ the constant of proportionality. Indeed, this
form of the field equations was first proposed by Einstein [2] and is remarkably close to the actual
Einstein field equations whilst being much simpler. The main problem lies with the divergence
of the tensor components, which is a sum over the covariant derivative Div(Dαβ ) = ∇α Dαβ .
It turns out the laws for the conservation of momentum and energy can be expressed in the
following form: ∇µ Tµν = 0. On the other hand, in general, ∇µ Rµν ̸= 0 therefore the proposed
form cannot hold.

However, when you add a correction term to form the Einstein tensor, Gµν = Rµν − 12 Rgµν ,
it turns out the divergence of this tensor is indeed zero, which can be shown using the Bianchi
identities (symmetries of the Riemann tensor). This yields the correct form of the Einstein
field equations, namely Gµν = κTµν . The constant of proportionality κ can be derived by
taking the Einstein field equations to the Newtonian limit in the case of weak gravity and low
velocities. The overview is as follows: firstly you assume the metric gµν = ηµν + hµν where η is
the Minkowski metric for flat space and h is a small perturbation. The low velocities condition
means that the velocity four-vector, U ⃗ has time component c and all other components close to
zero as u << c. The Einstein field equations then give you R00 = 12 κρc2 and using Poisson’s
i

equation this reduces to c12 ∇2 ϕ = 12 κρc2 finally yielding κ = 8πG


c4
= 2.077 × 10−43 s2 m−1 kg −1
meaning a lot of mass is needed to produce a noticeable amount of curvature.

The metric compatibility property states that ∇α gµν =0 meaning the Einstein field equations
can be still valid if you add a Λgµν term to them. This term is called the cosmological constant
and is used to represent dark energy driving the accelerated expansion of the universe. The
Einstein Field equations using the Friedmann-Lemaitre-Robertson-Walker/FLRW metric can
be used to derive the corrected Friedmann equations encoding the 3 large-scale geometries of
the universe.

13
5 The FLRW metric
The FLRW metric forms the basis for the standard model of cosmology and along with the field
equations, predicts an expanding universe. Recall that the cosmological principle stated that
the universe could be treated as being spatially homogeneous and isotropic; this means any
large-scale geometry of the universe must be consistent with the cosmological principle. There
are only 3 such geometries- flat, spherical and hyperbolic under the assumption that they are
simply connected i.e. have no holes like that of a torus.

5.1 The Universe as a Perfect Fluid


In the FLRW model, the matter and energy in the universe is modelled as a perfect fluid which
means it has a uniform mass density ρ and particles exert an equal amount of pressure p in
all directions. For flat space-time in Cartesian coordinates, the energy-momentum tensor for a
perfect fluid (2 raised indices) has the following form:
 2 
ρc 0 0 0
 0 p 0 0
[T µν ] = 
 0 0 p 0 .

0 0 0 p
However in general, the energy momentum tensor components are in terms of the metric
components as well as the 4-velocity, T µν = (ρ + cp2 )U µ U ν − pg µν or equivalently,
Tµν = (ρ+ cp2 )Uµ Uν −pgµν . This formula can be further simplified by taking the 4-velocity vector
to be at rest, i.e. U 0 = c with all other components being zero. However, if all the matter and
radiation is at rest for the FLRW metric, with respect to which frame is this? Furthermore,
in the FLRW metric which frame is used to measure the the time coordinate, t? According
to special and general relativity there should be no objective time coordinate to measure the
universe’s expansion as all inertial reference frames are equivalent. It turns out for both cases
there exists a special frame of reference known as the cosmic rest frame.

5.2 The Cosmic Rest Frame


As mentioned earlier, the cosmic rest frame can be considered as a frame of reference with
which the matter/radiation in the universe is at rest. When the Cosmic Microwave Background
Radiation (CMBR) is measured from the surface of the Earth, it turns out for half the surface
of the Earth it appears to be red-shifted and for the other half, it appears blue-shifted. This
occurs due to the Earth’s relative motion since part of the surface facing the CMBR appears
red-shifted and vice-versa for the part facing away.

The cosmic rest frame is then the frame of reference with respect to which there is no observed
Doppler shift in the CMBR radiation, i.e. the frequency looks the same in all directions.
Additionally, the cosmic rest frame is also the frame of reference through which, at the largest
scales ≈ 100 million ly the average velocity of all the mass/radiation is zero, as there is a net
outward shift in all directions due to the expansion of the universe.

14
Figure 10: CMBR Doppler shift viewed from the surface of the Earth [13]

5.3 The Friedmann equations with dark energy


Using the FLRW metric as well as the energy-momentum tensor for a perfect fluid, you can
derive the corrected Friedmann equations through the 00 component of the EFE’s and the
trace-reversed form (summing the EFE’s with the inverse metric). As a result, you obtain (1)
2
and (3) with the addition of a Λc3 term, however (2) is invariant. Firstly, you need to define
the density parameter, as a ratio of the the observed density to the critical density, Ωx = ρρxc .
3H 2
Here, the critical density today, ρc,0 = 8πG0 , is defined at the maximum density at which the
universe can continue to expand forever, or equivalently the density at which the universe halts
its expansion after an infinite period of time.
2 2
Also, you see the (1) can be rewritten as, H 2 = 8πG 3
Λc
(ρ + 8πG 3κc
− 8πGρa 2 ). Here, the first term

ρ is the density of the ordinary matter and radiation in the universe. This means it can be
split up into its two components, the matter/dust density ρm and the radiation density ρr .
Furthermore, it is convenient to express the matter and radiation density parameters in terms
of their classical solutions (see section 3.5). Firstly, for the matter density, you have ρm = ρam,0 3

where the constant of proportionality, ρm,0 is the matter density in the universe today (as earlier
for the critical density constant). Likewise, for the radiation density you have ρr = ρar,0 4 where

again ρr,0 is the radiation density today.


Λc2
The second term, 8πG can be interpreted as a new form of energy density, namely the dark
energy density of the universe, ρΛ . Since the cosmological constant is assumed to be indepen-
dent of time, so is the dark energy density, meaning it too is constant, i.e. ρΛ (t) = ρΛ,0 . The
final term is more subtle, but can be thought of as the density resulting from the curvature of
3κc2 ρκ,0
space time, − 8πGρa 2 = ρκ = a2 , which represents the solution for a curvature dominated uni-

verse. Note the change in sign, it means for a hyperbolic universe the spatial density parameter
is positive and vice-versa for a spherical universe. [14]

This yields the form: H 2 = 8πG3


( ρam,o
3 + ρar,0 ρκ,0
4 + ρΛ + a2 ). Dividing through by the Hubble

constant H02 and using the formula for the critical density parameter today gives the final form
2
in terms of the density parameters: H
H02
= Ωam,0 Ωr,0
3 + a4 +ΩΛ + a2
Ωκ,0
(7). Taking the time parameter
t to be the current age of the universe yields a(t0 ) = a0 = 1 and H = H0 which means that
Ωm,0 + Ωr,0 + Ωκ,0 + ΩΛ = 1 so the density parameters can be thought of as a relative measure
of the composition of the universe. For our own universe the density parameters have been

15
measured as ΩΛ ≈ 0.73, Ωm,0 ≈ 0.27 and Ωr,0 ≈ Ωκ,0 ≈ 0.

You may notice that there isn’t explicitly a term for the dark matter density parameter in
(7), however in the context of forming expansionary models the dark matter density is indeed
accounted for by the matter density, as both obey the relationship ρ ∝ a13 . This means that in
principle, dark matter can be treated as the missing/non-baryonic matter in the universe that
still obeys the same cosmological solution.

5.4 A comparison of the different solutions to the Friedmann equations

√ dominated universe you have Ωm,0 ≈ Ωr,0 ≈


Dark energy dominated universe: For a dark energy
da
Ωκ,0 ≈ 0 and therefore (7)√ reduces to dt = aH0 ΩΛ which is a separable differential equation
with the solution a(t) = e ΩΛ H0 t meaning a dark energy dominated universe will expand expo-
nentially, known as the Big Rip.

Dark energy and Matter/Dust: This is the solution of the Friedmann equation that is most
likely to be the fate of our universe, as you take Ωr,0 ≈ Ωκ,0 ≈ 0 which seemsqto agree with obser-
da Ωm,0
vational data. Then, the Friedmann equation gives the following: dt
= H0 Ω Λ a2 + a
. Once

Ωm,0 1 3H ΩΛ t 23
more, this is a separable differential equation with the solution: a(t) = ( ΩΛ ) 3 (sinh 0
2
) .
This solution can be thought of as a combination of the classical solutions for matter domi-
nated and dark-energy dominated universes as for small t (at the beginning of the universe) you
2
have a(t) ∝ t 3 just like the√matter dominated universe and for very large t, using the relation
x
sinh x ≈ e2 yields a(t) ∝ e ΩΛ H0 t which was the dark energy dominated solution from earlier.
This type of solution is known as the Big Freeze and represents an accurate long term solution
for our universe.

Matter and curvature: For this solution you assume that ΩΛ ≈ Ωr,0 ≈ 0, and repeating the

Figure 11: Graphical representation of the different solutions. Here, the vacuum universe is the dark-energy
dominated solution from earlier and the WMAP7 is the solution for our universe- with matter and dark energy.

−Ωm,0
process from earlier leads an implicit solution in terms of the parameter θ as a = 2Ωκ,0
(1−cos θ)

16
Ωm,0
and t = 3 (θ − sin θ). This is the equation of a cycloid, representing the fact that all
2H0 (−Ωκ,0 ) 2
such universes will end in a Big Crunch. Also for small values of theta you obtain the rela-
2
tion a(t) ∝ t 3 meaning that for a universe with matter and positive curvature (like that of a
sphere), matter will dominate at the start before curvature takes-over and causes the universe
to collapse in on itself. [10]

5.5 Limitations of the solutions


The models/solutions discussed here are classical, meaning that they are inaccurate for the
period of inflation; the first 10−32 s of the universe. To correct this, the inflation theory was
developed which postulated that from around 10−36 s to 10−32 s the universe expanded expo-
nentially from as small as 10−26 m to the order of a meter before slowing down to the rate of
expansion characterised by the Friedmann equations. The fundamental question is then, what
was the driving force for this expansion rate faster than the speed of light? According to the
theory the universe’s energy density was originally dominated by the cosmological constant
over the period of inflation which later decayed to produce the matter and radiation in the
universe today. Additionally, the source of the cosmological constant/dark energy also remains
unexplained; physicists usually put this down to the vacuum energy from quantum fluctuations,
however the predicted value is around 10124 times larger than the measured value, resulting in
one of the biggest problems in physics- the vacuum catastrophe. These types of contradictions
are usually cleared up using fine-tuning, i.e. assuming that there is a classical contribution
to the vacuum energy density cancelling out to 124 decimal places. Rather trivially, such a
magical correction is very controversial, as it leaves more questions to answer. [4]

6 Conclusion and Remarks


Cosmology is one of mankind’s oldest disciplines, and perhaps our grandest venture. Indeed,
the main goal of this project was to build up an elementary understanding of the field of
cosmology; to see how general relativity could predict an expanding universe. Towards the
start, the foundational principles of forming a cosmological model were laid out, and then us-
ing classical/Newtonian physics it was shown how the Friedmann equations were really just
a statement of energy conservation coupled with the first law of thermodynamics. Although
the exploration of general relativity was mathematically limited, the physically relevant fea-
tures were discussed, with a focus on building intuition rather than a formalised approach (see
Appendix). Finally, having developed an elementary understanding of general relativity, the
Friedmann equations with their relativistic corrections could be solved to form models of the
different possible universes.

17
7 Appendix
7.1 Vectors
General relativity is a geometric theory of gravity so the motion of objects is fundamentally
described in terms of vectors. The transformation rules for vectors also allow you to develop
the Lorentz transform, which is necessary for the theory of gravity to be relativistic i.e. agree
with the postulates of special relativity.

The common way to write vectors is using the Cartesian basis i,j,k, however in general since
the basis vectors are arbitrary, in the context of general relativity they are represented using e⃗i
where i runs over the number of dimensions in the space. Since basis vectors are arbitrary, it
is useful to know how the vector components and bases transform under a change of basis. In
general, the following formulas can be used for a change of basis: e⃗˜i = Fij e⃗j (for basis vectors)
and v˜i = Bij v j (vector components). Here, the superscripts v i are used the indicate vector
components (rather than powers), and the repeated j’s in each formula (1 superscript and 1
subscript) are used to indicate a sum over j, as per the Einstein summation convention. Also,
the coefficients Fij and Bij can be thought of as entries in a matrix F and B (called the forward
and backward transform) used to carry out the linear transformation (as a change of basis can
be thought of as one) such that F = B−1 .

7.2 Co-vectors
Co-vectors can be thought of as linear functions that take vector inputs and output real num-
bers. Or formally, a co-vector α : V → R. Similarly to vectors, they are necessary to define
certain objects in general relativity, and in general, any tensor can be thought of as a collection
of vector-co-vector pairs (see 5.1).
ej ) = δji where δji = 0 if j ̸= i and 1
Like vectors, co-vectors also have a basis ϵi such that ϵi (⃗
if j = i (Kronecker delta). Geometrically, co-vectors can be thought of as stacks whose output
on a vector is the number of lines it pierces. Transformation rules: Co-vectors can be geomet-

Figure 12: A co-vector W acting on a vector v [3]

rically thought of as being the ’opposite’ of a vector since they transform in the opposite way
to ordinary vectors. You have ϵ˜i = Bij ϵj and the forward transform for co-vector components
with summations over j as earlier.

18
7.3 The Tensor Product and the general definition of a tensor
The tensor product is a mathematical operation that makes it convenient to define tensors in
general. For vectors you see that their components transform using the backward transform,
hence they are called (1,0) tensors. Likewise, co-vector components transform using the forward
transform and so they are (0,1) tensors. Finally, the metric tensor is an example of a Bi-Linear
form which are (0,2) tensors as they transform using two forward transforms. This pattern
in general is true: so an (m, n) tensor can be thought of as a mathematical object whose
components transform using m backward transforms and n forward transforms. However, when
we considered vectors and co-vectors it was natural to form a basis for these, and in general to
form a basis for a tensor you use what is called the tensor/outer product, ⊗. To demonstrate
this for the metric tensor suppose you have a basis ϵij , then you must have the following
identity, ϵij (e⃗µ , e⃗ν ) = δµi δνj where you have the product of Kronecker deltas. However, using the
properties of the tensor product, you can show that co-vector bases, ϵi ⊗ ϵj (e⃗µ , e⃗ν ) = δµi δνj . This
means that ϵi ⊗ ϵj form a basis for the metric tensor, or g = gij ϵi ⊗ ϵj .
This can also be generalised to form another equivalent definition of an (m, n) tensor. Namely
if T is an (m, n) tensor then: T = Tba11ba22...b ...am
n
(e⃗a1 ⊗ e⃗a2 ... ⊗ ϵb1 ⊗ ϵb2 ...). Equivalently, an (m, n)
tensor is a linear function acting on m co-vectors and n vectors, producing a real number (or
the component of the tensor).

7.4 Derivation of the geodesic equation


The geodesic equation which forms a key aspect of general relativity can be derived as follows.
Firstly, basis vectors can be given an equivalent definition in the context of multi-variable calcu-

∂X ⃗ 0
lus, e⃗i = ∂x i , where X represents a general 4-D space-time vector, with X = ct or t depending

on the choice of units.

Now, following the definition of a geodesic you need to consider the acceleration vector:

d2 X ⃗
d dxi ∂ X
dT 2
= dT ( dT ∂xi ) employing the multi-variable chain rule. Then, using the product rules along
d j ∂ 2⃗ ⃗
2 i ∂X ⃗
dxj dxi ∂ X
with the fact that dT = dx
dT ∂xj
, gives ddTX2 = ddTx2 ∂x i + dT dT ∂xi ∂xj . Using the tangential basis
⃗ ⃗
vectors and normal vectors to the 4-D manifold, you have ∂x∂iX ∂xj
∂X
= Γkij ∂x k + Lij n̂(*). Here, the

Γkij represent the Christoffel symbols, which form the components of the tangential vectors and
the Lij represents the normal components. Finally, substituting this in and setting the tangen-
2 k j dxi
tial components to zero gives the geodesic equation: ddTx2 + Γkij dx dT dT
= 0 (4). The fact that the
tangential components are zero follows from the definition of a geodesic and the equivalence
principle (all inertial frames are indistinguishable and follow the same physical laws).

7.5 The Covariant derivative


The covariant derivative is a mathematical tool that allows you to track changes in a vector field
along a given direction or path parameter. Formally, the covariant derivative ∇w⃗ ⃗v measures the
change in the vector field ⃗v along the direction of w.
⃗ It is also defined to be linear, additive and
satisfies the Leibniz product rule. The covariant derivative can then be defined by how it acts
on the basis vectors, or by computing ∇e⃗i e⃗j . Since the covariant derivative of a basis vector
along another basis vector is itself a vector, you have ∇e⃗i e⃗j = Γkij e⃗k . Here, the Γkij are denoted
connection coefficients (thought of as a generalisation of the Christoffel symbols). However in
the context of general relativity it is desirable for the Christoffel symbols to be unique, so two
more properties can be added to the definition of the covariant derivative. Firstly, we say that
the covariant derivative is torsion-free, meaning that ∇⃗v w ⃗ = ∇w⃗ ⃗v or equivalently Γkij = Γkji .
Secondly, the metric compatibility property means that ∇⃗u (⃗v · w) ⃗ = ∇⃗u (⃗v ) · w ⃗ + ∇⃗u (w)
⃗ · ⃗v .

19
The concept of the covariant derivative is closely associated with the idea of parallel trans-
porting a vector, which itself can be thought of as a way to transport a vector ⃗v along a curve
w(λ)
⃗ such that it remains as ’straight as possible’. This can be formalised using the covariant
derivative as follows; a vector ⃗v is parallel transported along w(λ)
⃗ if ∇w(λ)
⃗ ⃗ v = 0. Note that if
two vectors are parallel transported, by the metric compatibility property the angle between
them is preserved.

Figure 13: Parallel transporting a vector on a sphere along RQP [7]

Using the concept of parallel transport, a geodesic is then a curve which can be parallel
⃗ = 0 ⇔ (4).
transported along itself, ∇X⃗ X

7.6 Riemann Tensor components


This Riemann curvature tensor can be formalised using the covariant derivative as follows,
⃗ ⃗t = ∇⃗v ∇w⃗ ⃗t − ∇w⃗ ∇⃗v⃗t − ∇[⃗v,w]
R(⃗v , w) ⃗
⃗ t. Where taking the covariant derivative along w ⃗ then ⃗v
corresponds to travelling along w ⃗ then ⃗v and vice versa for the second term. The final term
[⃗v , w]
⃗ is the Lie bracket and can be thought of as the error term in a curved geometry where in
general the ends of the parallelogram don’t meet.
The components of the Riemann tensor can be defined by how it acts on basis vectors, or
R(e⃗µ , e⃗ν )e⃗σ = Rσµν
ρ
e⃗ρ = ∇µ ∇ν e⃗σ − ∇ν ∇µ e⃗σ − ∇[µ,ν] e⃗σ . Since the Lie bracket for basis vectors
ρ
goes to zero, you have: Rµν e⃗ρ = ∇µ Γkνσ e⃗k − ∇ν Γkµσ e⃗k = ∂µ Γkνσ e⃗k + Γkνσ Γlµk e⃗l − ∂µ Γkµσ e⃗k − Γlνk Γkµσ e⃗l
(note the change in notation for the partial and covariant derivatives: ∂µ = ∂X∂ µ and ∇e⃗µ = ∇µ )
. Re-indexing gives Rσµν ρ
= ∂µ Γνσ + Γkνσ Γρµk − ∂ν Γρµσ − Γkµσ Γρνk .

7.7 Calculating the Christoffel symbols


The Christoffel symbols are calculated as follows: ∇e⃗µ (⃗ ei · e⃗j ) = ∇e⃗µ (⃗
ei ) · e⃗j + ∇e⃗µ (⃗
ej ) · e⃗i . Using
metric compatibility you have ∇e⃗µ gij = Γkµi e⃗k · e⃗j + Γkµj e⃗k · e⃗i . But as the covariant derivative
∂g
of a scalar is just the partial derivative, you have ∂µ gij = ∂Xijµ = Γkµi gkj + Γkµj gki (i) (note the
change in the notation for the partial derivative). Likewise upon permuting µ, i and j you also
have ∂j gµi = Γkjµ gki + Γkji gkµ (ii) and ∂i gjµ = Γkij gkµ + Γkiµ gkj (iii). Doing i+ii-iii and using the
torsion-free property yields 2Γkµj gki = ∂µ gij + ∂j gµi − ∂i gjµ . Finally, using the inverse metric (a
(2,0) tensor) you get Γkµj g im gki = 12 g im (∂µ gij + ∂j gµi − ∂i gjµ ) and using the index cancellation
1 im
law g im gki = δkm yields Γm µj = 2 g (∂µ gij + ∂j gµi − ∂i gjµ ) (5). This property uniquely defines
the Christoffel symbols through the metric tensor, meaning knowing the metric for a space you
have the Christoffel symbols.

20
7.8 Proof of the divergence-free property of the Einstein tensor
Firstly, in order to simplify the formulas you have the following notation ∇σ Xab = Xab;σ . Then,
ρ ρ ρ
the second Bianchi identity states that Rσµν;α + Rσαµ;ν + Rσνα;µ = 0 ⇔ Rρσµν;α + Rρσαµ;ν +
Rρσνα;µ = 0 (lowering the Riemann tensor components using the metric).

Next, you multiply and sum through using two inverse metric tensors, g σν g ρµ (Rρσµν;α +
Rρσαµ;ν + Rρσνα;µ ) = 0. Distributing the first term and raising the components you have
g σν (Rσµν;α
µ µ
+ Rσαµ;ν + g ρµ Rρσνα;µ ) = 0. Using the 3-4 and 1-2 symmetries of the Riemann
tensor, you have g σν (Rσµν;α
µ µ
− Rσµα;ν − g ρµ Rσρνα;µ ) = 0. You see that the 1st and 2nd terms are
really just Ricci tensor components which gives g σν (Rσν;α − Rσα;ν − g ρµ Rσρνα;µ ) = 0. Finally,
ν ν
distributing the 2nd inverse metric and raising the indices gives Rν;α − Rα;ν − g ρµ Rρνα;µ
ν
= 0.
As the first term is just the Ricci scalar and the other two terms are the raised index versions
of the Ricci tensor, you have R;α − Rα;ν ν µ
− Rα;µ = 0. Re-indexing gives 12 R;α − Rα;ν
ν
. Multiplying
1 µν µν
through by the inverse metric once more gives ( 2 Rg − R );ν = 0. Lowering the indices you
finally have Gµν;ν = 0, as needed to show that the Einstein Tensor is divergence-free.

7.9 The metric tensor for flat space


Since the universe can be thought of as expanding radially outwards, it is more convenient to
represent the metric tensor in polar coordinates. Furthermore, since the basis vectors have to
encode the changing distances/ changing scale of space-time you have e⃗i → a(t)⃗ei . This yields
the metric as follows:
 
1 0 0 0
0 −(a(t))2 0 0 
g= 2 2

0 0 −r (a(t)) 0 
2 2 2
0 0 0 −r sin θ(a(t))
.

7.10 The metric tensor for a closed/spherical space


Remembering the Cartesian equation of a circle/S 1 as x2 + y 2 = R2 , this can be generalised to
the n-sphere/S n as x21 + x22 + · · · + x2n+1 = R2 . For the FLRW metric, the universe is treated
spatially as a 3-sphere/S 3 meaning you have x2 + y 2 + z 2 + w2 = R2 . Similar to the case for
flat space, it is more convenient to work in spherical coordinates where you use the 3 circular
angles
 θ, ϕ and χ. The metric for spherical space then has  the form:
1 0 0 0
0 −(a(t))2 0 0 
g= 2 2
.
0 0 − sin χ(a(t)) 0 
2 2 2
0 0 0 − sin χ sin θ(a(t))

7.11 The metric for hyperbolic/open space


Analogous to the n-sphere S n , the generalisation of the hyperbola defined by x2 − y 2 = R2
to n-dimensions is called H n defined as x21 − x22 − · · · − x2n+1 = R2 , or the hyperbolic plane.
This is perhaps the most interesting geometry since the metric tensor for a hyperboloid H 2 in
Cartesian coordinates is not the identity matrix, and therefore actual distances on the surface
defy physical intuition. This is also illustrated by the Poincare disc model (shown below) where
each triangle shown has the same area. .

Continuing the pattern as earlier the universe’s spatial geometry can be represented by H 3
in terms of hyperbolic angles θ, ϕ and χ. This finally yields the metric:

21
Figure 14: The Poincare disc model for H 2 where lines are represented as arcs of circles [16]

 
1 0 0 0
0 −(a(t))2 0 0 
g= 2 2
.
0 0 − sinh χ(a(t)) 0 
2 2 2
0 0 0 − sinh χ sin θ(a(t))

7.12 Unifying the 3 metrics


There are a lot of similarities in the 3 metrics for the flat, spherical and hyperbolic geometries
of the universe. In fact, they can be combined into a single metric by introducing the curvature
term (as discussed earlier) k = 1, 0, −1 and introducing a change of coordinates χ → r as
follows: r = χ for a flat universe, r = sin χ for a spherical universe and r = sinh χ for a
hyperbolic universe. Furthermore as now the metric must be represented in the (t, r, ϕ, θ) co-
ordinate system you need to calculate grr which using the chain rule is grr = ( dr1 )2 gχχ . For a

dr 2
flat universe ( dχ ) = cos2 χ = 1 − sin2 χ = 1 − r2 and finally for
dr 2
) = 1, for a closed universe ( dχ
an open universe ( dχ ) = cosh2 χ = 1 + sinh2 χ = 1 + r2 . Using the k parameter these formulas
dr 2

dr 2
can be combined as ( dχ ) = 1 − kr2 . This yields the final form of the general FLRW metric as:
 
1 0 0 0
0 −(a(t))22 0 0 
g= 1−kr
2 2
.
0 0 −r (a(t)) 0 
0 0 0 −r2 sin2 θ(a(t))2

References
[1] behrouz. Hubble’s Law - Astro Data Science. Sept. 2021. url: https : / / astrodatascience . net /
hubbles-law/.
[2] Albert Einstein. Volume 6: The berlin years: Writings, 1914-1917 (english translation supplement) page
103. 1915. url: https://einsteinpapers.press.princeton.edu/vol6-trans/115.
[3] Marcelo Epstein. A physicist’s view of a covector ω and its evaluation on a Vector V. 2023. url: https:
//www.researchgate.net/figure/A-physicists-view-of-a-covector-o-and-its-evaluation-on-
a-vector-v_fig1_323989967.
[4] Guth. Inflationary cosmology. 2023. url: https://ned.ipac.caltech.edu/level5/March05/Guth/
Guth1.html.
[5] James B Hartle. Gravity: an introduction to Einstein’s general relativity. 2003.
[6] R Nave. Kepler’s Laws. url: http://hyperphysics.phy-astr.gsu.edu/hbase/kepler.html.
[7] J Pan. Information geometry. May 2009. url: https : / / gdpan919 . wordpress . com / 2009 / 05 / 15 /
information-geometry/.

22
[8] M. Pettini. Institute of Astronomy. 2023. url: https://www.ast.cam.ac.uk/students/undergrad/
part_ii/lectures/cosmology.
[9] Dmitri Pogosyan. CMBR. 2023. url: https://sites.ualberta.ca/~pogosyan/teaching/ASTRO_122/
lect31/lecture31.html.
[10] Aula Raul. 2.4 solutions to the friedmann equation - university of São Paulo. url: http://fma.if.usp.
br/~mlima/AulaRaul2018/Scale-Factor.pdf.
[11] Erhard Scholz. “From heliocentrism to epicycles: A commentary on pre-Ptolemaic astronomy”. In: arXiv
preprint arXiv:2208.02137 (2022).
[12] James Schombert. Geometry of the Universe. 2013. url: http : / / abyss . uoregon . edu / ~js / cosmo /
lectures/lec15.html.
[13] telescoper. A cosmic microwave background dipole puzzle. Oct. 2016. url: https://telescoper.wordpress.
com/2016/10/31/a-cosmic-microwave-background-dipole-puzzle/.
[14] tutorialspoint. Cosmology - hubble amp; density parameter. July 2021. url: https://www.tutorialspoint.
com/cosmology/cosmology_hubble_and_density_parameter.htm.
[15] Hans Uy. The mystery of Mercury’s missing arcseconds. May 2020. url: https://www.astronomicalreturns.
com/2020/05/the-mystery-of-mercurys-missing.html.
[16] Eric Weisstein. Poincaré hyperbolic disk. 2004. url: https://mathworld.wolfram.com/PoincareHyperbolicDisk.
html.

23

You might also like