Spacetime Compactification in General Relativity

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 36

The University of Nottingham

Physics Research Project


F34PJM

Spacetime Compactification
in General Relativity

Author: Supervisor:
Thomas Goulding Dr. Anastasios Avgoustidis

May 29, 2015


Abstract. The aim of this project is to investigate the properties of different space-
times by visualising them. These spacetimes are the solutions to Einstein’s general theory
of relativity. Using Penrose-Carter diagrams, an infinite four-dimensional universe can be
depicted in a finite plot. This technique is utilised in order to study the differences between
various general relativistic solutions. Conformal transformations allow the universe to be
mapped on a finite scale, without losing information about the geometry of the spacetime.
Plotting Penrose-Carter diagrams illustrates the key properties of the solutions, showing the
past and future of the universe. As well as creating these diagrams, an example of their
use in modern research is discussed, which involves reproducing the results from a recently
released paper. This research is then furthered in this report by introducing new features
that are hitherto unexplored.

Contents
1 Introduction 1

2 Background Physics 1
2.1 Minkowski Spacetime . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
2.2 De Sitter Spacetime . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2.3 Anti-de Sitter Spacetime . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.4 Robertson-Walker Spacetime . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.5 Schwarzschild Spacetime . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.6 The Kerr Solution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.7 Taub-NUT Spacetime . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

3 Method and Results 10


3.1 Minkowski Spacetime . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
3.2 De Sitter Spacetime . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
3.3 Anti-de Sitter Spacetime . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
3.4 Robertson-Walker Spacetime . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.4.1 Robertson-Walker with k = +1 . . . . . . . . . . . . . . . . . . . . . . . . 17
3.4.2 Robertson-Walker with k = 0 . . . . . . . . . . . . . . . . . . . . . . . . . 18
3.4.3 Robertson-Walker with k = −1 . . . . . . . . . . . . . . . . . . . . . . . . 20
3.5 Schwarzschild Spacetime . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
3.6 Topological Defects from the Multiverse . . . . . . . . . . . . . . . . . . . . . . . 24
3.6.1 Loops with α = 0 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
3.6.2 Loops with α 6= 0 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

4 Discussion 31
4.1 Minkowski Spacetime . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
4.2 De Sitter and Anti-de Sitter Spacetimes . . . . . . . . . . . . . . . . . . . . . . . 31
4.3 Robertson-Walker Spacetime . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
4.4 Schwarzschild Spacetime . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
4.5 Topological Defects from the Multiverse . . . . . . . . . . . . . . . . . . . . . . . 32

5 Conclusion 33

6 References 33
1 Introduction
Penrose-Carter diagrams, also called Penrose diagrams or conformal diagrams, are illustrations of
a spacetime, plotted in terms of time, η, and space, χ. They are used to map potentially infinite
spacetimes in finite size. Each spacetime will be represented by a Penrose-Carter diagram of a
different shape, often with different boundaries. Large, even infinite, values for space and time
are fit onto the diagram by changing the coordinate system used, thus shrinking distances that
are further away. This distorts the geodesics that represent constant space and time, bending
them from their usual straight shape to produce curves.
In a region of spacetime, any two events can be connected by joining them with a group of
curves. The curve in this group with either the maximum or minimum proper length is known
as a geodesic [Griffiths and Podolský, 2009].
Experimental and theoretical exploration of the curved spacetimes of general relativity is
conducted by studying the motion of particles and light through them. A "test" body is used,
which is a particle with mass small enough to cause no significant spacetime curvature. The
motion of this body is determined by the curvature produced by other, more massive, bodies.
The paths followed by test bodies through a curved spacetime are called "timelike" geodesics
[Hartle, 2002], and they represent constant space. Similarly, the paths taken by massless particles
(light rays) are the geodesics for light, usually referred to as "null" geodesics, and they are
always represented on a Penrose-Carter diagram as straight lines at angles of ±45◦ [Mukhanov,
2005]. While the timelike paths that extremise proper time are known as timelike geodesics,
the spacelike paths that minimise the proper distance between two points are called "spacelike"
geodesics [Blau, 2014].
In this report, these three types of geodesics will be studied in the case of several different
spacetimes. The spacetimes covered here are by no means every discovered model. Rather, the
solutions which are the most interesting or the most important to cosmological research will be
investigated.

2 Background Physics
This section of the report will cover the background physics required to understand the subject
material. Previous examples from other publications will be studied, which will be useful when
later discussing the results obtained from this project.

2.1 Minkowski Spacetime


Minkowski spacetime is an empty model used in special relativity. Mathematically, it is a four-
dimensional flat Euclidean space, denoted by R4 . The metric of a spacetime defines its properties,
for example curvature and chronology. In this case, the metric is given in Cartesian coordinates
by:
ds2 = −c2 dt2 + dx2 + dy 2 + dz 2
And in spherical polar coordinates by:

ds2 = −c2 dt2 + dr2 + r2 dθ2 + sin2 θdφ2




where r ∈ [0, ∞], θ ∈ [0, π], φ ∈ [0, 2π].

1
Fig. 1. The Penrose-Carter diagram for Minkowski spacetime [Mukhanov, 2005].

After performing conformal transformations, this spacetime can be illustrated by a finite


Penrose-Carter diagram, as shown in Figure 1. In this diagram, the new timelike coordinate (η)
goes in the vertical direction and space (χ) is represented horizontally. Therefore, the bottom
corner shows the past (i− ) and the top corner shows the distant and infinite future (i+ ). These
are the start and end points, respectively, of timelike geodesics. Similarly, the corner on the right
represents spacelike infinity i0 . Light rays follow the null geodesics at ±45◦ from the bottom-


left edge (I − ) to the top-right edge (I + ). Hence the boundary of Minkowski spacetime consists
of null infinity (I ± ), timelike infinity (i± ) and spacelike infinity i0 [Frauendiener, 2004]. The


curved lines represent constant space, r, and constant time, t. These are the variables from the
spacetime before compactification. The shaded region I is the future light cone - the area which
a flash of light radiating in all directions would cover [Mukhanov, 2005]. The timelike infinity,
i+ , is a single point in the compactified Minkowski spacetime and it is the apex of the shaded
light cone at infinity [Jadczyk, 2010].

2.2 De Sitter Spacetime


While Minkowski spacetime is flat, the de Sitter universe has a positive four dimensional cur-
vature (R > 0), although it remains homogeneous and isotropic. This particular spacetime is
commonly used to study inflation, although a de Sitter universe with a slightly broken time-
translational symmetry must be used for it to be relevant to physical applications [Mukhanov,
2005].

The metric for de Sitter space is given by:

cosh2 (HΛ t) sin2 r


    
k = +1
ds2 = −dt2 + HΛ−2  exp (2HΛ t)  dr2 +  r2  dΩ2  k=0
sinh2 (HΛ t) sinh2 r k = −1
1
where HΛ = (8πGεΛ /3) 2 and dΩ2 = dθ2 + sin2 θdφ2 .


The value k corresponds to the curvature of the spacetime, with a closed universe having k = +1,
a flat universe having k = 0, and an open universe having k = −1.

2
Fig. 2. The conformal diagram for de Sitter spacetime [adapted from Mukhanov, 2005].

As seen in Figure 2, the Penrose-Carter diagram for the compactified de Sitter universe is a
simple square shape, with two diagonal horizons. These are the past and future horizons of a
person at the right or left edge - a classical observer sitting on the right edge can never observe
anything past the diagonal line that connects the bottom left of the diagram to the top right, and
they can only ever send messages to the part of space above the diagonal line from bottom right
to top left [Spradlin et al., 2001]. This is illustrated in Figure 3. In Figure 3 (i), the shaded area
shows the space that can be observed by someone on the right-hand edge of the diagram, while
(ii) illustrates the part of space which this observer can send messages to. These are referred
to as the causal past and future of the observer respectively. Null geodesics originate from the
bottom of the diagram (I − ) and terminate at the top (I + ), with light rays travelling at ±45◦
similar to the Penrose-Carter diagram for Minkowski spacetime.

(i) (ii)

Fig. 3. Causal past and future diagrams.

2.3 Anti-de Sitter Spacetime


The homogeneous solution with constant negative curvature (R < 0) is known as anti-de Sitter
spacetime, and in terms of curvature it is the "opposite" to de Sitter spacetime [Sun and Tian,
2014]. It is a symmetric solution to Einstein’s general theory of relativity, with a negative cosmo-
logical constant - this means that it has an attractive effect on the universe. This is a property
which is not physically observed in our universe, but it corresponds to an extra attractive term
in the gravitational force [Bengtsson, 1998].

The metric for anti-de Sitter spacetime is given by:

ds2 = −cosh2 rdt2 + dr2 + sinh2 r dθ2 + sin2 θdφ2




Before constructing the Penrose-Carter diagram for this solution, the r coordinate is rescaled to
shrink down the values of infinity.

3
Fig. 4. Penrose diagram of anti-de Sitter spacetime [Hawking and Ellis, 1973].

Figure 4 shows the Penrose-Carter diagram for this spacetime. The boundary of the timelike
surface I makes up infinity, along with the two disjoint temporal infinities - i− for the past
infinity and i+ for the future infinity. Also shown is the null geodesic for a light ray from a
point labelled "p" to the timelike surface. The points p and q are the spots at which all timelike
geodesics converge, with p in the past and q in the future [Bengtsson, 2011]. All timelike geodesics
converge to a point before diverging to refocus at a second point, and this repeats. This means
that the timelike geodesics never reach the timelike surface I , unlike the future null geodesics.
The points p and q lie on the line representing conformal space χ = 0, here referred to as r0 = 0.
χ has a maximum value of π2 which coincides with I [Hawking and Ellis, 1973].

2.4 Robertson-Walker Spacetime


Known as the Standard Model of modern cosmology, the Robertson-Walker spacetime is more
general than the de Sitter model, in that it allows for forms of energy such as matter and ra-
diation, rather than just vacuum energy [Bergström and Goobar, 2008]. Deriving the Einstein
field equations in this spacetime leads to the Friedmann equations, which are the equations that
govern the expansion of space in a homogeneous and isotropic universe and describe how the
scale factor of our universe evolves [Nemiroff and Patla, 2008]. The full name that is sometimes
used for this spacetime is the Friedmann-Lemaître-Robertson-Walker (FLRW) model.

The Robertson-Walker metric is given by:

ds2 = −c2 dt2 + a2 (t) dr2 + Sk2 (r)dΩ2


 

where
 
r
Sk (r) = R0 sin , k = +1
R0
= r, k=0
 
r
= R0 sinh , k = −1
R0

and dΩ2 = dθ2 + sin2 θdφ2 . a(t) is the scale factor of the universe, and R0 is the present-day


curvature radius. At any time, the curvature radius is given by a(t)R0 , but at a time t = t0 (the
present), a(t) = a(t0 ) = 1.
The appearance of the Penrose-Carter diagram for Robertson-Walker spacetime depends on
the value of the cosmological constant (Λ). This is the value of the vacuum energy density of
space [Solà, 2013].

4
Fig. 5. Penrose-Carter diagrams for Robertson-Walker spacetimes with vanishing cosmological constant,
Λ = 0 [Griffiths and Podolský, 2009].

The first two Penrose diagrams shown in Figure 5 correspond to Robertson-Walker spacetimes
with curvature k = +1 and no cosmological constant. The first of these two diagrams represents a
spacetime which contains dust, and the second illustrates a spacetime which contains radiation.
The wavy lines represent singularities - the universe expands from the big bang singularity,
reaches a maximum size, and then re-collapses to a final singularity. The dashed lines denote
the path taken by a photon emitted from χ = 0 soon after the big bang. In the radiation-filled
universe the photon only reaches χ = π, but the dust-filled universe lasts longer so the photon
encircles the whole space and returns to its starting point [Griffiths and Podolský, 2009].
The Penrose diagram in the centre of Figure 5 corresponds to a flat FLRW spacetime (k = 0).
There is again the big bang singularity at η = 0, by which the universe is bounded, as well as
the future infinity I + . The point i+ and i0 denote timelike and spacelike infinities respectively.
The final Penrose-Carter diagram in Figure 5, on the right, shows an open universe (k = −1).
It is very similar to the flat universe Penrose diagram, bounded by a big bang singularity and
future infinity I + , as well as having future timelike (i+ ) and spacelike i0 infinities. However,
the coordinate lines that correspond to constant space and time are not tangent to the axes near
i0 or i+ , whereas in the flat universe the geodesics are tangent to the η = 0 singularity near the
spacelike infinity, and tangent to the χ = 0 line close to the timelike infinity.

Fig. 6. Penrose-Carter diagrams for Robertson-Walker spacetimes with positive cosmological constant,
Λ > 0 [Griffiths and Podolský, 2009].

In Figure 6, the diagram on the left shows the Penrose-Carter diagram for a Robertson-
Walker spacetime with Λ > 0 and k = +1, the middle diagram shows the FLRW universe
with positive cosmological constant but for a flat universe, and the rightmost diagram shows an
open spacetime (k = −1). For all three universes, the spacetimes begin at an initial big bang
singularity and continue to expand to an infinite size. The positive cosmological constant causes
this infinite expansion to occur very rapidly, such that the future conformal infinity I + occurs
in a finite time η. This means that I + is spacelike.

5
Fig. 7. Penrose-Carter diagrams for Robertson-Walker spacetimes with negative cosmological constant,
Λ < 0 [Griffiths and Podolský, 2009].

The leftmost Penrose diagram in Figure 7 shows a closed (k = +1) FLRW spacetime, the
centre diagram is for a flat universe, and the furthest on the right shows a spacetime with k = −1.
All of these diagrams correspond to a universe with a negative cosmological constant. All three
universes are bounded by an initial big bang singularity at η = 0 and a final singularity at a
finite conformal time η > 0. All of these spacetimes expand to a maximum size, shown by the
dotted lines on the diagrams, before collapsing to the final spacelike singularity.

2.5 Schwarzschild Spacetime


The Schwarzschild solution represents the spherically symmetric empty spacetime outside a
spherically symmetric massive body [Hawking and Ellis, 1973]. Therefore, it is used in describing
the local geometry of spacetime in the solar system, and is widely applied in considerations of
orbital motions about the Sun or the Earth [Griffiths and Podolský, 2009]. This solution to
Einstein’s field equations was found only a few months after the equations were published. Until
recently, all experiments carried out to test the differences between general relativity and New-
tonian theory were based on this solution. This spacetime shows the theory of relativity to be
superior to Newton’s classical theory of gravitation, by predicting the small differences between
Newtonian theory and physical observations. For example orbital motion in the solar system,
or deflection of light by the Sun, as well as gravitational time delay [Griffiths and Podolský, 2009].

The metric for the Schwarzschild solution is given by:

2m −1 2
   
2 2m 2
dr + r2 dθ2 + sin2 θdφ2

ds = − 1 − dt + 1 −
r r
where m is an arbitrary parameter which is usually interpreted as representing the mass of a
body.

Fig. 8. Penrose diagram for the Schwarzschild spacetime [Griffiths and Podolský, 2009].

6
Figure 8 illustrates the complete structure of the Schwarzschild spacetime, where all the lines
shown represent constant r. As discussed previously, the Schwarzschild spacetime provides an
accurate representation of the gravitational field around a spherically symmetric massive object,
such as a star. Towards the end of the life of a star, if it is massive enough, its density can
become large enough that its outer surface approaches the critical value of r = 2m. Eventually
it may even collapse to form a black hole. Particles (timelike or null trajectories) can pass in
through r = 2m but they cannot come out, so r = 2m is known as a horizon, and the spacetime
within the horizon is the black hole. No event that occurs inside this horizon can ever be seen
by an outside observer so it is referred to as an "event horizon".
Einstein’s field equations have no preferred time direction, and therefore there is another
singularity which is a time-reversed black hole, known as a "white hole", from which light can
emerge but none can reach. It is referred to as a white hole because the singularity can be seen.
Nothing from outside r = 2m can enter through this horizon in an inward direction. Rather
than being called an event horizon, the horizon around a white hole is more commonly referred
to as a "Cauchy horizon". It should be noted that no real physical situation like this has yet
been observed. In addition, because the spacetime does not extend beyond the singularity, this
means that nothing could produce it [Griffiths and Podolský, 2009].
In Figure 8, the lines of constant r approach the spacelike infinities i0 , the timelike infinities i−
and i+ , and the "central" point. The Cauchy horizon for the white hole is a spacelike surface near
the initial big bang-like singularity, and the event horizon for the black hole is shown above the
white hole horizon in the diagram. Outside these two horizons, there are two separated regions
for r > 2m which are very similar to Minkowski spacetime, which shows that Schwarzschild
spacetime is asymptotically flat [Carroll, 1997]. The different sections of the Schwarzschild
Penrose diagram are emphasised in Figure 9.

Fig. 9. Diagram to illustrate the separation of the black hole, the white hole, and the asymptotically
Minkowski-like spacetime.

2.6 The Kerr Solution


Most astronomical bodies are rotating and therefore the spacetime outside them is not expected
to be spherically symmetric [Hawking and Ellis, 1973]. As explained in the previous section,
the exterior of a spherically symmetric body is represented by the Schwarzschild spacetime.
However, a rotating body can collapse to form a black hole and due to the conservation of
angular momentum, the rate of rotation will increase. The end result of a stationary rotating
black hole is an important situation to model. A model that describes a stationary, axially
symmetric, asymptotically flat spacetime is required and in the case of a rotating black hole, an
event horizon must also be included. The spacetime that meets all of these requirements is the
Kerr solution [Griffiths and Podolský, 2009].

7
The metric for the Kerr solution can be expressed as:

∆r 2 %2 2 sin2 θ  2
ds2 = − 2
dt − a sin2
θdφ + dr + %2 2
dθ + 2
adt − r2 + a2 dφ
% ∆r %

where %2 = r2 +a2 cos2 θ and ∆r = r2 −2mr +a2 . m is the mass of the source and a is interpreted
as a rotation parameter. When a = 0 the metric reduces to a standard spherical Schwarzschild
solution.

Fig. 10. The conformal diagram for the Kerr spacetime with constant θ and φ
[adapted from Griffiths and Podolský, 2009].

Figure 10 shows the Penrose-Carter diagram for the Kerr solution, and it can be seen that for
a region such as the one labelled I, the horizons with r = r+ have the character of a black hole
event horizon - events inside the horizon cannot be seen by an observer at i+ . r = r− has the
character of a Cauchy horizon. The shaded areas correspond to "ergoregions", which are regions
of the Kerr spacetime just outside the black hole, inside which nothing can be seen [Griffiths and
Podolský, 2009]. Within these regions it is not possible to remain at rest with respect to infinity
but it is still possible to escape to infinity [Hawking and Ellis, 1973].

2.7 Taub-NUT Spacetime


The Taub-NUT spacetime takes its name from the discoverers Taub, Newman, Unti and Tam-
burino. Initially expressed by Taub as the time-dependent part of the spacetime, it was redis-
covered by the three other scientists as a simple generalisation of the Schwarzschild spacetime,
this time covering both stationary and time-dependent regions. The Taub-NUT solution is a
simple non-radiating solution of Einstein’s field equations, although it has been argued that this
spacetime has no physical interpretation [Griffiths and Podolský, 2009]. However, it is still an
interesting solution to consider.

8
The metric for the complete Taub-NUT spacetime is a complex equation:

4l2
 
2 2 2 4l 2 1 2 2
 41 2
ds = F −dT± + dR± − (R± dT± − T± dR± ) sin θdφ + 2 T± − R± sin θdφ
r± 2 r± 2

+ r 2 + l2 dθ2 + sin2 θdφ2


 

where: 1−r∓ /r±


3r

4r± ∓ r − r∓
F = 2 e−r/r±
(r + l2 ) r∓

r − r+ 1/2 r − r− r− /2r+ r/2r+


     
t
T+ = e sinh
r+ r− 2r+
 1/2  r− /2r+  
r − r+ r − r− r/2r+ t
R+ = e cosh
r+ r− 2r+
 r+ /2r−  1/2  
r+ − r r− − r r/2r− t
T− = e sinh
r+ r− 2r−
 r+ /2r−  1/2  
r+ − r r− − r t
R− = er/2r− cosh
r+ r− 2r−

l is the NUT parameter, which is interpreted as the "magnetic mass" or "gravitomagnetic


monopole moment".

Fig. 11. The conformal diagram for the Taub-NUT spacetime [Griffiths and Podolský, 2009].

The conformal diagram for this spacetime, illustrated in Figure 11, shows lines of constant
r. In this diagram, the values of θ and φ are constant. The complete Taub-NUT spacetime
is made up of an infinite sequence of Taub regions separated by alternating NUT+ and NUT−
regions. These two regions correspond to r > r+ for NUT+ and r < r− for NUT− . The Taub
regions separating these areas relate to r− < r < r+ , in which the coordinate r is timelike and t
is spacelike [Griffiths and Podolský, 2009].

9
3 Method and Results
This section of the report will cover the mathematical method for producing the Penrose diagrams
from the initial spacetime metrics. It will also include the conformal diagrams produced with
these methods.

3.1 Minkowski Spacetime


The metric for the empty Minkowski universe is given by:
ds2 = −c2 dt2 + dr2 + r2 dΩ2
where dΩ2 = dθ2 + sin2 θdφ2 .


The first step in compactifying this metric is to introduce null coordinates given by:
u=t−r
v =t+r
The infinities in this spacetime can be limited to a finite range by applying the following trans-
formation:
p = arctan(u) = arctan (t − r)
q = arctan(v) = arctan (t + r)
Which then leads to timelike and spacelike coordinates defined by:
η = p + q = arctan (t − r) + arctan (t + r)
χ = q − p = arctan (t + r) − arctan (t − r)
The ranges of the coordinates are now limited to η ∈ [−π, π] and χ ∈ [0, π], giving a new metric:
ds̃2 = −dη 2 + dχ2 + sin2 χdΩ2
The coordinates η and χ can be plotted against each other for constant time (spacelike) and
constant space (timelike), to show the geodesics on a Penrose-Cater diagram. This is shown in
Figure 12.

Fig. 12. The Penrose diagram for Minkowski spacetime. Spacelike geodesics are shown as pink lines
converging at the spacelike infinity on the right of the diagram. Timelike geodesics are illustrated as red
lines that converge at the negative and positive timelike infinities (the top and bottom corners). The null
geodesics are plotted as straight lines at 45◦ .

10
So far, the angular coordinates θ and φ have been suppressed, but these coordinates can now
be reintroduced to the metric and added as a feature on the diagram. Leaving φ as a constant,
meaning that dφ = 0, the value of θ can be varied from 0 to 2π. In the figure above, each
point on the diagram represents a 2-sphere of radius sin χ. Therefore, when one of the angular
coordinates is reintroduced, the resulting three-dimensional shape has a radius of zero at 0, π,
and -π. This can be seen in the conformal diagram in Figure 13 (i). In order to make the Penrose
diagram easier to interpret, a false radius will be used - the 2-spheres will have a radius of χ
instead of the real value of sin χ. The diagram using this false radius of rotation is shown in
Figure 13 (ii).

(i) (ii)

Fig. 13. (i) This is the conformal diagram for the Minkowski spacetime with angular coordinate dθ
reintroduced, showing the past and future conformal infinities, which have maximum values at η = −π/2
and η = π/2 respectively. (ii) The conformal diagram for Minkowski spacetime, with a modified 2-sphere
radius of χ representing the true value of sin χ. The null geodesics, shown here in blue, have now become
null cones.

Both angular coordinates cannot be introduced together because it is impossible to produce


a four-dimensional diagram. However, suppressing dθ by making the value of θ constant, dφ can
be reintroduced. This has the same effect as when dθ was introduced, and again the modified 2-
sphere radius will be used to improve the clarity of the diagram. In this case however, the radius
of rotation has a value of sin χ sin θ (or with modified radius, χ sin θ), where θ is a constant which
can be varied. The conformal diagrams produced when this angular coordinate is reintroduced
are shown in Figure 14.

Fig. 14. Four conformal diagrams for Minkowski spacetime with angular coordinate dφ introduced, using
the modified radius χ sin θ. Each diagram shows the compactification with different values of θ, from 0.1
to π/2. At θ = π/2, the diagram is identical to that in Figure 13 (ii).

11
3.2 De Sitter Spacetime
For the de Sitter space, the metric for curvature k = +1 will be used, and this is given by:

ds2 = −dt2 + cosh2 t dr2 + sin2 rdΩ2




where dΩ2 = dθ2 + sin2 θdφ2 .




In the method for the Minkowski spacetime, new spacelike and timelike coordinates were intro-
duced, and constant values for space and time were used to plot timelike and spacelike geodesics
respectively. In this method for the de Sitter metric, however, the Euler-Lagrange equations
will be used. This is a different method for finding the geodesics of a spacetime. The first
step towards defining the geodesic equation for de Sitter space is to find an expression for the
Lagrangian (ignoring the angular coordinates for now):
 2
ds
L= = −ṫ2 + cosh2 t ṙ2

d
where ˙ ≡ dλ .

This can be used in the Euler-Lagrange equation for r:


   
dL d dL
− =0
dr dλ dṙ
 
dL
= cosh2 t ṙ = k
dṙ
For geodesics, the Lagrangian is constant:

L = const.

L cosh2 t = k 2 − cosh2 t ṫ2


A new variable v can be introduced, set as:

v = tanh t

And by differentiating, it can be shown that:

v 02 = M 2 − v 2

where M 2 = 1 + L/k 2 .

Some timelike geodesics can be found by observing:

k = 0 ⇒ r = const.

But in the case where k 6= 0, the differential equation must be solved. This solution is given by
an equation relating t and r, and this is the equation for the geodesics of this spacetime:

tanh t = M sin r

In the case of timelike geodesics L = 1, for null geodesics L = 0, and for spacelike geodesics
L = −1. The value of M is dependent on the value of L, and therefore M has different ranges for
each type of geodesic. By plotting the equation above for the correct values of M , the geodesics
for the de Sitter spacetime can be illustrated. These are the non-compactified geodesics but they
can be plotted on a diagram, shown in Figure 15.

12
Fig. 15. Diagram for part of de Sitter spacetime in original coordinates before compactification. Shown
here are the timelike geodesics in red, null geodesics in blue, and spacelike geodesics in pink.

To produce the Penrose diagram for this spacetime, a new timelike coordinate is used:

dη 2 = cosh−2 t dt2

The new timelike coordinate is related to the original timelike coordinate by:

η = arctan (sinh t)

This compactifies infinite time t to η = π/2, and r can be limited by introducing χ = r for values
of r between 0 and π. Therefore the metric now has a new form:

ds̃2 = cosh2 t −dη 2 + dχ2 + sin2 χdΩ2




Plotting the new timelike and spacelike coordinates against each other generates the Penrose-
Carter diagram, shown in Figure 16 (i).

(i) (ii)

Fig. 16. (i) The Penrose diagram for de Sitter spacetime with compactified geodesics, using the new
timelike and spacelike coordinates. The green lines represent the timelike geodesics in the case where
k = 0. The red lines show timelike geodesics for k 6= 0, the pink lines show the spacelike geodesics, and
the blue lines at 45◦ show the null geodesics. (ii) The conformal diagram for de Sitter spacetime with
angular coordinate dθ reintroduced, using a false 2-sphere radius of cosh η sin χ.

The angular coordinates can now be introduced, starting with dθ and suppressing dφ. From
the equation for the line element, it can be seen that the radii of the 2-spheres at each point in

13
the Penrose diagram are given by cosh t sin χ. Putting this in terms of the new coordinates, the
radius is given by cosh (arcsinh (tan η)) sin χ, but to improve the clarity of the diagram, a false
radius of rotation can be used, as in the case for the Minkowski spacetime. In this case, the false
radius used will be cosh η sin χ, which is effectively a compactified radius. Therefore, taking each
2-sphere to have this new radius, the angular coordinate dθ can be reintroduced to produce a
new conformal diagram, shown in Figure 16 (ii).

Fig. 17. Four conformal diagrams for de Sitter spacetime with angular coordinate dφ reintroduced. The
false rotation radius used here is cosh η sin χ sin θ, and each diagram uses a different value of θ from 0.1
to π/2.

The same process can now be repeated when reintroducing the other angular coordinate, dφ.
This time, the coordinate dθ is suppressed by setting θ to have a constant value. Again, the
compactified 2-sphere radius of cosh η sin χ will be used. In this case, the full expression for the
false rotation radius is given by cosh η sin χ sin θ and as a result, the size of the diagram can be
varied by changing the value of constant θ. The conformal diagrams produced when this angular
coordinate is reintroduced are shown in Figure 17.

3.3 Anti-de Sitter Spacetime


The anti-de Sitter metric is given by:

ds2 = −cosh2 rdt2 + dr2 + sinh2 rdΩ2

where dΩ2 = dθ2 + sin2 θdφ2 .




The geodesic equation for this spacetime is found using the Euler-Lagrange equations, simi-
lar to the method used for de Sitter spacetime, rather than simply defining new spacelike and
timelike coordinates. The first step in this process is to define the Lagrangian for the non-angular
coordinates:  2
ds
L= = − cosh2 r ṫ2 + ṙ2

d
where ˙ ≡ dλ .

Similar to the method for de Sitter spacetime, the Euler-Lagrange equation for t is then used:
   
dL d dL
− =0
dt dλ dṫ
 
dL
= cosh2 r ṫ = k
dṫ

14
It is known that the Lagrangian is constant for geodesics:

L = const.

L cosh2 r = cosh2 r ṙ2 − k 2


A new variable dependent upon r can now be introduced:

u = tanh r

Differentiating, it is found that:


u02 = M 2 − u2
where M 2 = 1 − L/k 2 .

One class of spacelike geodesics is easily found:

k = 0 ⇒ t = const.

For the case where k 6= 0, to find the geodesic equation, the differential equation must be solved.
The solution in this case is given by:

tanh r = M sin t

The Lagrangian has different constant values for each type of geodesic (L = 1 for timelike, L = 0
for null, and L = −1 for spacelike). Using the same method as for the de Sitter spacetime,
these different values of L change the value of M in the geodesic equation. A diagram showing
the non-compactified geodesics of anti-de Sitter spacetime can be constructed by plotting the
equation above for different values of M . This diagram is shown in Figure 18.

Fig. 18. Diagram of the non-compactified geodesics for part of anti-de Sitter spacetime. Timelike geodesics
are shown as red lines, null as blue, and spacelike as pink.

Compactification of this spacetime is performed by defining a new spacelike coordinate:

dχ2 = cosh−2 r dr2

Hence the new spacelike coordinate can be related to the original coordinate by:

χ = arctan (sinh r)

Consequently, infinite space is compactified to χ = π/2. The range of the timelike coordinate
can be restricted between 0 and π by introducing a new timelike coordinate η = t, with limits.
Therefore the metric now has a new form:

ds̃2 = cosh2 r −dη 2 + dχ2 + tan2 χdΩ2




15
The new timelike and spacelike coordinates can now be plotted against each other to produce
the Penrose diagram for anti-de Sitter spacetime, shown in Figure 19 (i).

(i) (ii)

Fig. 19. (i) Penrose diagram for anti-de Sitter spacetime with the geodesics now compactified to finite
space. Timelike geodesics are shown in red, spacelike in pink, and null in blue. The green lines represent
the spacelike geodesics for the case where k = 0. (ii) Conformal diagram for anti-de Sitter spacetime with
angular coordinate dθ reintroduced, using a 2-sphere radius of χ rather than the true value of tan χ.

Like the method for the previous spacetimes, the angular coordinate dθ can now be reintro-
duced while suppressing dφ. It can be seen from the metric that the radius of each 2-sphere in
the Penrose diagram for anti-de Sitter spacetime is given by tan χ. Using this value when intro-
ducing the angular coordinates makes the diagram very unclear. To make the diagram easier to
interpret, a false 2-sphere radius will again be used. In this case, the radius will be given a value
of χ, which means that the Penrose diagram will simply be rotated around χ = 0 (Figure 19
(ii)).
If the angular coordinate dθ is now suppressed, the coordinate dφ can instead be introduced.
This has the same effect as before, but with an additional factor of sin θ in the expression for the
2-sphere radius. A false radius can again be used for clarity, which in this case will have a value
of χ sin θ, where the true value would be tan χ sin θ. The constant θ can therefore be varied and
this will change the size of the diagram, shown in Figure 20.

Fig. 20. Conformal diagrams for anti-de Sitter spacetime. These diagrams represent the spacetime with
angular coordinate dφ introduced, using a false 2-sphere radius of χ sin θ. Each of the four diagrams uses
a different value of θ between 0.1 and π/2.

16
3.4 Robertson-Walker Spacetime
The Friedmann-Lemaître-Robertson-Walker cosmological Standard Model has a metric given by:

ds2 = −c2 dt2 + a2 (t) dr2 + Sk2 (r)dΩ2


 

where
 
r
Sk (r) = R0 sin , k = +1
R0
= r, k=0
 
r
= R0 sinh , k = −1
R0

and dΩ2 = dθ2 + sin2 θdφ2 . a(t) is the scale factor of the universe, which has a present-day


value of 1. R0 is the present-day curvature radius which can be ignored due to its constant
nature, and c is the speed of light which can be reduced to 1 using a geometric unit system.
In this part of the report, only the Robertson-Walker spacetime with vanishing cosmological
constant (Λ = 0) will be investigated.

3.4.1 Robertson-Walker with k = +1


Compactification of the FLRW spacetime for k = +1 is not necessary, as it is already conformally
related to the Einstein static universe - a universe infinite in both space and time that is neither
expanding nor contracting. In this case, r = χ and t = η, with both coordinates having a range
between 0 and 2π.
ds2 = −dη 2 + dχ2 + sin2 χdΩ2
where dΩ2 = dθ2 + sin2 θdφ2 .


No compactification is performed on this metric, and as a result, the spacelike and timelike
geodesics are represented by straight lines in the Penrose diagram (Figure 21 (i)).

(i) (ii)

Fig. 21. (i) The Penrose diagram for FLRW spacetime with k = +1, with straight timelike geodesics
(shown in red, vertical) and spacelike geodesics (pink, horizontal). Null geodesics are plotted at 45◦
(blue). (ii) Conformal diagram for Robertson-Walker spacetime with k = +1 and reintroduced angular
coordinate dθ.

17
Reintroducing the angular coordinate dθ is a fairly simple process for this space, with each
point in Figure 21 (i) corresponding to a 2-sphere of radius sin χ. With φ set to be a constant,
such that dφ is zero, a conformal diagram can be produced, which is shown in Figure 21 (ii).
The diagrams produced for compression of dθ (and introduction of dφ) look very similar to
Figure 21 (ii), but this time have a rotation radius of sin χ sin θ. By using a different value of θ
in each diagram, the radius is changed by a factor of sin θ. The conformal diagrams produced
when reintroducing this angular coordinate are shown in Figure 22.

Fig. 22. Conformal diagrams for FLRW spacetime with k = +1 and reintroduced angular coordinate dφ,
for different values of constant θ.

3.4.2 Robertson-Walker with k = 0


In the case where k = 0, compactification is necessary due to both coordinates (r and t) ranging
from 0 to ∞. This metric is conformally related to half of Minkowski spacetime, and is given by:

ds2 = −dt2 + dr2 + r2 dΩ2

where dΩ2 = dθ2 + sin2 θdφ2 .




Compactification of this spacetime begins by introducing a new set of coordinates:

u=t−r
v =t+r

These can then be conformally transformed by setting:

p = arctan(u) = arctan (t − r)
q = arctan(v) = arctan (t + r)

Finally, new timelike and spacelike coordinates can be defined:

η = p + q = arctan (t − r) + arctan (t + r)
χ = q − p = arctan (t + r) − arctan (t − r)

These coordinates lead to a new metric given by the expression:

ds̃2 = −dη 2 + dχ2 + sin2 χdΩ2

This is equivalent to part of the Einstein static universe, as in the case where k = +1, but
with η ∈ [0, π] and χ ∈ [0, π]. A Penrose-Carter diagram can be produced for this spacetime by

18
plotting the new coordinates against each other, while keeping t constant for spacelike geodesics
and keeping r constant for timelike geodesics (Figure 23 (i)). Timelike geodesics in this diagram
converge at the future timelike infinity (top corner), and spacelike geodesics converge at the
spacelike infinity (right corner). These geodesics become tangent to the axes close to their
respective infinities.

(i) (ii)

Fig. 23. (i) Penrose diagram for FLRW spacetime with k = 0 and Λ = 0. Timelike geodesics appear here
as red lines and spacelike geodesics are illustrated as pink lines. Null geodesics are again blue lines at 45◦ .
(ii) The conformal diagram of the FLRW universe with k = 0. This shows the compactified geodesics
as well as the reintroduced angular coordinate dθ, using a false radius of rotation χ in place of the true
value of sin χ.

The angular coordinate dθ can be reintroduced, while setting dφ to zero. Similar to the
method used with Minkowski spacetime, a false radius will be used. It is clear from the metric
for this universe that in the figure above, each point on the diagram should represent a 2-sphere
of radius sin χ. However, to improve the clarity of the conformal diagram, a false radius of χ
will be used for rotation instead of the real value. This will therefore result in a simple rotation
around χ = 0, and produce the diagram shown in Figure 23 (ii).
The diagrams for the FLRW spacetime for k = 0 with dφ introduced and dθ suppressed can
be constructed in a similar way. Again, the 2-sphere radius used is χ rather than the real value,
sin χ. The diagrams are similar to Figure 23 (ii), but with a factor of sin θ changing the rotation
radius. This can be seen in Figure 24.

Fig. 24. Conformal diagrams for FLRW spacetime with k = 0, and reintroduced angular coordinate dφ,
using a false 2-sphere radius of χ. Each diagram uses a different value for θ, between 0.1 and π/2.

19
3.4.3 Robertson-Walker with k = −1
The third and final FLRW spacetime which will be investigated here is the case where k = −1.
The metric for this universe is given by:

ds2 = −dt2 + dr2 + sinh2 rdΩ2

where dΩ2 = dθ2 + sin2 θdφ2 .




A new coordinate set can be defined as:

u=t−r
v =t+r

These coordinates can then be compactified by setting:


 
u t−r
p = tanh = tanh
2 2
v   
t+r
q = tanh = tanh
2 2

These p and q coordinates can be altered by introducing:


  
t−r
p̃ = arctan(p) = arctan tanh
2
  
t+r
q̃ = arctan(q) = arctan tanh
2

Finally, new timelike and spacelike coordinates can be given by:


     
t+r t−r
η = q̃ + p̃ = arctan tanh + arctan tanh
2 2
     
t+r t−r
χ = q̃ − p̃ = arctan tanh − arctan tanh
2 2

This leads to a new expression for the metric that is the same form as the line elements of the
previous two Robertson-Walker spacetimes, meaning that it is again equivalent to part of the
Einstein static universe:
ds̃2 = −dη 2 + dχ2 + sin2 χdΩ2
A Penrose diagram can be produced by plotting the new timelike and spacelike coordinates
against each other. Timelike geodesics are plotted on the graph using constant values for space,
and spacelike geodesic lines are plotted by holding values of time constant (Figure 25 (i)). This is
very similar to the diagram for FLRW spacetime with k = 0, with timelike geodesics converging
at the timelike infinity (top corner), and spacelike geodesics again converging at the spacelike
infinity (right corner). The difference to notice in this case is that the geodesics do not become
tangent to the axes.

20
(i) (ii)

Fig. 25. (i) Penrose-Carter diagram for Robertson-Walker spacetime with negative curvature, k = −1.
Timelike geodesics are represented as red lines, and spacelike geodesics are represented as pink lines. Null
geodesics are shown as straight blue lines. (ii) Conformal diagram for FLRW spacetime with k = −1,
and angular coordinate dθ introduced. This diagram shows a rotation of sin χ.

As is clear from the equation for the line element of this spacetime, the radius of the 2-
spheres on each point on the Penrose diagram is sin χ. Therefore when reintroducing the angular
coordinate dθ (keeping φ constant), the result will be a rotation of radius sin χ, seen in Figure 25
(ii). The result of the sine function applied to χ is that the spacelike infinity has been moved
from π/2 to 1, while the timelike coordinates remain unchanged.
The process is similar when reintroducing the angular coordinate dφ. This time, the value of
θ must remain constant, but with a factor of sin θ affecting the radius of rotation, the rotation
can be performed for different values of constant θ. The conformal diagrams with the angular
coordinate dφ reintroduced is shown in Figure 26.

Fig. 26. Conformal diagrams with angular coordinate dφ, for the FLRW spacetime with k = −1. In each
of the four diagrams, a different value of θ is used (between 0.1 and π/2). In all of the diagrams, the
rotation radius has a value of sin θ sin χ.

21
3.5 Schwarzschild Spacetime
The metric for the Schwarzschild solution is given by:

2m −1 2
   
2 2m 2
ds = − 1 − dt + 1 − dr + r2 dΩ2
r r

where dΩ2 = dθ2 + sin2 θdφ2 , and m is an arbitrary parameter which is usually interpreted as


mass.

To begin the process of compactification, a variable called the "tortoise coordinate" is intro-
duced, defined as: r
r∗ = r + 2Gm ln − 1

2Gm
This is then incorporated into a new set of null coordinates, given by:

u = t − r∗
v = t + r∗

These are Eddington-Finkelstein coordinates for a Schwarzschild spacetime and they are called
the "retarded null coordinate" and "advanced null coordinate" respectively. However, this coor-
dinate system is still badly behaved in some parts of the spacetime, in particular where r = 2m.
This issue can be resolved by reparameterising these null coordinates by introducing:

−u
  r 1  r  
−t

2
U = − exp =− − 1 exp exp
4m 2m 4m 4m
 v   r 1  r  
t

2
V = exp = − 1 exp exp
4m 2m 4m 4m
These are then combined to produce new timelike and spacelike coordinates defined by:
1 1
T = (V + U ) , R= (V − U ) , for r > 2m
2 2
1 1
T = (V − U ) , R = (V + U ) , for r < 2m
2 2
These new coordinates can be related to the original Schwarzschild coordinates by:
 r 1  r   
t 
2
T = − 1 exp sinh 
2m 4m 4m 
 r 1  r    r > 2m
2 t 
R= − 1 exp cosh 

2m 4m 4m
r  21  r   
 t 
T = 1− exp cosh 
2m 4m 4m 
r < 2m
r  12  r   
 t 
R= 1− exp sinh 

2m 4m 4m

These are known as Kruskal-Szekeres coordinates. Although this metric has not yet been com-
pactified, a diagram can be constructed to show the timelike and spacelike geodesics (Figure 27).
Some parts of the diagram have been cut off in this figure, due to the fact that the full range
extends to R ∈ [−∞, ∞] and T ∈ [−∞, ∞] at the horizons.

22
Fig. 27. The diagram for the Schwarzschild spacetime in Kruskal-Szekeres coordinates with θ and φ
constant. Spacelike geodesics are shown in pink, timelike geodesics are shown in red. Null geodesics
appear as straight blue lines at 45◦ .

To compactify this spacetime fully, and produce the conformal diagram, the null coordinates
from the Kruskal-Szekeres coordinate system are rescaled and given by:

Ũ = 2 arctan U
Ṽ = 2 arctan V

Which leads to the timelike and spacelike coordinates:


1  1 
η= Ṽ + Ũ , χ= Ṽ − Ũ , for r > 2m
2 2
1  1 
η= Ṽ − Ũ , χ= Ṽ + Ũ , for r < 2m
2 2
Which can then be plotted to produce the Penrose diagram for Schwarzschild spacetime, shown
in Figure 28 (i). In this diagram, the horizons can be clearly seen along the null geodesic lines,
shown in blue. Also seen in this figure are the two asymptotically Minkowski-like regions for
r > 2m. These two exterior regions are connected by a section which is behind horizons and
has a minimum value for r at R = 0. This is known as an "Einstein-Rosen bridge", or more
commonly referred to as a "wormhole". No real particles can pass through this from one exterior
region to the other [Griffiths and Podolský, 2009].
Looking back at the Schwarzschild metric, the angular coordinates can be reintroduced,
starting with θ. From the metric, it is clear that each point on the conformal diagram in
Figure 28 (i) represents a 2-sphere with radius r. Taking φ to be constant so that dφ = 0, and
with dθ 6= 0, a three-dimensional conformal diagram is produced, shown in Figure 28 (ii).

23
(i) (ii)

Fig. 28. (i) The Penrose diagram for the complete Schwarzschild spacetime with θ and φ constant,
showing spacelike, timelike and null geodesics as well as the infinities. (ii) The conformal diagram for
Schwarzschild spacetime with the angular coordinate dθ reintroduced.

This method can be repeated when reintroducing angular coordinate dφ, such that φ is not
constant, with different values of θ, to produce the conformal diagrams shown in Figure 29. This
time, the rotation radius in the diagram is multiplied by a factor of sin θ.

Fig. 29. Conformal diagrams for Schwarzschild spacetime with angular coordinate φ reintroduced, for
four different values of θ.

3.6 Topological Defects from the Multiverse


This part of the report will focus on the use of Penrose diagrams in modern research. Of
course, compactification can be useful in analysing theoretical models, and one example of this is
investigating topological defects in the multiverse. Specifically, these defects are loops nucleating
in the parent vacuum outside of the bubble of our universe. Using a de Sitter slicing, the
propagation of these defects can be analysed.
A string loop defect which nucleates in the parent vacuum of the multiverse will be oriented
by an angle α. The case where α = 0 has been previously studied by Zhang et al. [2015], and
this will be reproduced in this report and then researched further by considering a loop with an
angle of inclination α 6= 0.

3.6.1 Loops with α = 0


Firstly the nucleation of a loop at an angle of α = 0, such that it is parallel to the x-y plane, can
be examined. The spacetime outside of our universe is known as the "false vacuum", and is a de
Sitter space with an expansion rate HF . The spacetime inside our bubble is the "true vacuum"
and can be approximated by a de Sitter spacetime with an expansion rate of HT ≤ HF .

24
The worldsheet of our bubble is approximated its light cone, defined by the equation:

r = 1 − e−HF t

where r(t = 0) = 0.

The worldsheet for a loop nucleated at a point x = rn and a time t = tn can be described
by the equation:
(x − rn )2 + y 2 + z 2 = e−2HF tn + e−2HF t
where z = (x − rn ) tan α.

The future light cone of the loop approaches the worldsheet of the loop as t → ∞. Figure 30 (i)
shows all potential loops, with the loop of α = 0 highlighted.

(i) (ii)

Fig. 30. (i) All potential loops are illustrated here as a sphere, with the chosen loop (α = 0) shown as a
blue circle. (ii) Diagram showing the worldsheet of the loop with α = 0 in blue, and the worldsheet of a
bubble in orange, with the line of intersection illustrated in red.

The worldsheet for this chosen loop can then be constructed, along with the light cone of our
bubble, which nucleates at a different time and space. The collision of the two worldsheets can
be seen in Figure 30 (ii).
The nucleation centre coordinates can be redefined in terms of the conformal coordinates in
closed de Sitter slicing:
sin ηn
e−HF tn = −
cos ηn + cos χn

sin χn
rn =
cos ηn + cos χn

For the nucleation centre to be in the spacetime covered by these coordinates, the following
condition must be met:
χn ≤ η n + π

Focussing on the x-axis only, where y = z = 0, the nearest and most distant points on the loop
can be found:
s
sin χn sin2 ηn
xL± = ± e−2HF t +
cos ηn + cos χn (cos ηn + cos χn )2

xB± = ± 1 − e−HF t


where the subscript L indicates the coordinates for the loop, and B denotes the coordinates for
the bubble. ± shows the right and left parts respectively.

25
When analysing a collision between the bubble and a defect, we will assume that the nucle-
ation centre of the loop is always on the right of the bubble nucleation centre (rn > 0).
For any collision to occur between the loop and bubble, we must have the right side of the
loop inside the left side of the bubble, or the left side of the loop inside the right side of the
bubble, at late times:
xL+ < xB− or xL− > xB+ at t → ∞
As t → ∞, e−HF t → 0, which means the above condition can be written as:

|xL− (t → ∞)| < 1


s
2

sin χn sin ηn
cos ηn + cos χn − (cos ηn + cos χn )2 < 1

−π/2 < χn + ηn < π/2

For the loop to completely enter the bubble, another condition must be satisfied. The side of
the loop not considered by the previous condition must also be inside the bubble. This means
that the right side of the loop must be inside the right side of the bubble, and the left side of
the loop must be inside the left side of the bubble, also at late times:

xL+ < xB+ or xL− > xB− at t → ∞

Since only the right side of the bubble is being considered, the first of these two options is all
that is needed:

xL+ (t → ∞) < 1
s
sin χn sin2 ηn
+ <1
cos ηn + cos χn (cos ηn + cos χn )2

χn − ηn < π/2

To find the collisions which begin on the right of the bubble, a third and final condition is needed.
This is simply that the loop originates on the right side of the bubble (x > 0). If the left side of
the loop is on the right side of the bubble, then the entirety of the loop can be assumed to be
on that side:

xL− (t = 0) ≥ 0
s
sin χn sin2 ηn
− 1+ <1
cos ηn + cos χn (cos ηn + cos χn )2

If a collision occurs where the loop does not fully enter the bubble, and the condition for a
collision on the right is not satisfied, then the collision can be assumed to start on the left side
of the bubble.
These inequality regions can now be represented on a Penrose-Carter diagram for de Sitter
spacetime, as seen in Figure 31. The nucleation of the bubble occurs at χn = 0 and ηn = −π/2.
Any loops that have nucleation centres inside the bubble will completely enter the bubble and
this area is represented in red. Loops can also form at a time before bubble nucleation in a region
such that, when the bubble nucleates, it will initially be encircled by the loop, but eventually
the bubble will expand and fully contain the loop at late times. Meanwhile, any loops with
nucleation centres in the green area will encircle the bubble but only partially enter it on the
left side. Any loops with nucleation centres in the blue area of the diagram will enter the bubble
from the right.

26
Fig. 31. Penrose diagram for de Sitter spacetime, showing the regions which satisfy different conditions
for a collision to occur between the loop and the bubble. Loops with nucleation centres in the red area
will fully enter the bubble. Loops with nucleation centres in the green and blue regions will only partially
enter the bubble, either from the left or the right side respectively.

3.6.2 Loops with α 6= 0


The results from Zhang et al. [2015] can be expanded upon by introducing a value of α 6= 0.
Again, the expression for a loop nucleating at (tn , rn ) is given by:

(x − rn )2 + y 2 + z 2 = e−2HF tn + e−2HF t

where z = (x − rn ) tan α.

(i) (ii) (iii)

Fig. 32. (i) A diagram showing the sphere of all potential loops, with the chosen loop at α = 0.7
highlighted. The red line shows the radius of the previous loop with α = 0, and the green line shows the
size along the x-axis of the new loop. (ii) The same diagram viewed from above, at 90◦ to the x-y plane,
for clarity. (iii) Diagram to show the worldsheet of the projected ellipse for the loop with angle α = 0.7,
and the light cone of the bubble, with the line of intersection between them shown here in red.

Figure 32 (i) shows a loop with α = 0.7 chosen from the potential loops. The green and
red lines show a comparison between the x-direction size of this loop and the x-direction size of
the previously investigated loop with α = 0. There are many ways that the new loop can be
interpreted for analysis, but in this report, a projection onto the x-y plane will be used. This
produces an ellipse from the angled loop, and this can be seen in Figure 32 (ii), when the diagram
is viewed from above.

27
The worldsheet for the projected ellipse can now be constructed, and the light cone for a
bubble is introduced. These are both shown in Figure 32 (iii). The line where these two world-
sheets intersect is highlighted in red.

The radius of the chosen angled loop has a radius given by:

R = ((x − rn )2 + y 2 + z 2 )1/2 = (e−2HF tn + e−2HF t )1/2

where, in spherical coordinates:

x = rn + R cos θ sin φ
y = R sin θ sin φ
z = R cos θ

However, the value of z is now redefined as a function of the angle of inclination, α:

z = (x − rn ) tan α

This means that the size along the x-axis of the ellipse projected by the loop can be given by:
1/2
x = R2 − y 2 − z 2 + rn
t 1/2
1/2
= e−2HF tn + e−2HF 1 − sin2 θ sin2 φ − cos2 θ sin2 φ tan2 α

+ rn

The spacelike coordinate of the nucleation centre can again be defined in terms of conformal
coordinates, as in the case for α = 0, and this value is neither changed by the introduction of
the angle, nor by the projection of the loop when producing the ellipse:
sin χn
rn =
cos ηn + cos χn
Therefore the value of rn can be taken out of the expression for the x-direction size of the ellipse
for now. Since this analysis will be focussing on only the x-axis with y = z = 0, this can
be expressed in spherical coordinates as θ = 0 and φ = π/2, which reduces the x-coordinate
expression to:

x = ((e−2HF tn + e−2HF t ) − (e−2HF tn + e−2HF t ) tan2 α)1/2


= (e−2HF tn + e−2HF t )1/2 (1 − tan2 α)1/2

The timelike coordinate of the loop nucleation centre can also be written in terms of conformal
coordinates, and this has the same form as the timelike coordinate for a loop with inclination
angle α = 0, as this remains unchanged by the angle and projection:
sin ηn
e−HF tn = −
cos ηn + cos χn
The bubble remains unchanged by any of the above processes. Therefore, the equation for the
points on the bubble remains the same. The equations for the nearest and most distant points
on the loop and bubble are now given by:
s
2 
sin χn sin η n
xL± = ± e−2HF t + (1 − tan2 α)
cos ηn + cos χn (cos ηn + cos χn )2

xB± = ± 1 − e−HF t


Again, the subscripts L and B denote the coordinates for the loop and bubble respectively. ±
differentiates the right and left parts.

28
The conditions for collisions that were used in the previous section can be applied again, us-
ing the modified equation for the loop with α 6= 0. For any collision to occur, the left side of the
loop must be inside the right side of the bubble, or the right side of the loop must enter the left
side of the bubble, at late times:

xL+ < xB− or xL− > xB+ at t → ∞

As t → ∞, e−HF t → 0, which means the above condition can be written as:

|xL− (t → ∞)| < 1


s
2

sin χn sin ηn 2

cos ηn + cos χn − (cos ηn + cos χn )2 (1 − tan α) < 1

For a collision to occur such that the entire loop is inside the bubble at late times, the other side
of the loop must also enter the bubble. This means that the right side of the loop must be inside
the right side of the bubble, or the left side of the loop must enter the left side of the bubble:

xL+ < xB+ or xL− > xB− at t → ∞

Only the right side of the bubble is being considered here, so the first of these conditions is used:

xL+ (t → ∞) < 1
s
sin χn sin2 ηn
+ (1 − tan2 α) < 1
cos ηn + cos χn (cos ηn + cos χn )2

Finally, the condition must be applied for collisions occurring on the right. A loop which partially
enters the bubble will enter on the right if the loop nucleates with its left side on the right of the
bubble. This can be expressed by the condition:

xL− (t = 0) ≥ 0
s
sin2 ηn

sin χn
− 1+ (1 − tan2 α) ≥ 0
cos ηn + cos χn (cos ηn + cos χn )2

As in the case for α = 0, the regions which satisfy each of these conditions can now be represented
on a Penrose-Carter diagram of de Sitter spacetime. This process can be performed for a range
of values of α, from 0 to π/2. Figure 33 shows a sample of these diagrams. As before, loops with
nucleation centres in the red region will fully enter the bubble. Loops which have nucleation
centres in the blue or green parts of the spacetime will only partially enter the bubble, either
from the right or left respectively.
As alpha is increased above 0, the blue region of the Penrose diagram begins to dominate
over the green region. The red area of the spacetime also grows, until a value of α = π/4,
where the minimum timelike coordinate of this region reaches negative timelike infinity. For
π/4 ≤ α < π/2, the red area totally dominates, with the green and blue regions no longer
visible. At α = π/2, all areas disappear.
Arguably the most interesting value of α is π/4, the angle at which the green and blue areas
representing a partial collision vanish. Figure 34 shows the conformal diagram for this spacetime
with α < π/4, but only marginally less than this value. It can be seen in this diagram that
the green region is almost non-existent. Although the blue region is dominating in the areas
satisfying the partial collision conditions, it too is becoming vanishingly small.

29
Fig. 33. Four conformal diagrams of de Sitter spacetime, each showing the regions for different types of
collisions between a loop and the bubble, for a range of values of angle of inclination α (between π/10
and 2π/5).

Fig. 34. Penrose diagram for de Sitter spacetime showing the regions satisfying the bubble-defect collision
conditions. This diagram is for a loop at an angle α = π/4 − 0.01.

In order to investigate how these different regions vary with the angle of the loop, a graph
can be plotted to illustrate the proportion of the spacetime covered by each region as a function
of α (Figure 35).
From this graph it is clear that the green region of the de Sitter spacetime, which satisfies
the conditions for a collision to occur on the left of the bubble, decreases at an increasing rate.

30
The blue region, which represents the area which satisfies the conditions for a partial collision
to occur on the right side, increases slowly and then rapidly vanishes. Both of the regions for
partial collisions vanish at α ≈ π/4. The red region, satisfying the conditions for the loop to
fully enter inside the bubble, grows rapidly until the value of α reaches ∼ π/4. Then, for angles
greater than this value, the red region decreases in size until it disappears at α = π/2.

Fig. 35. A graph showing the fraction of the de Sitter spacetime covered by each of the regions satisfying
the collision conditions, plotted against the loop orientation angle, α.

4 Discussion
This section of the report will be used for discussion of the results obtained in this project. By
looking back at the diagrams produced using the methods described in the previous section, they
can be compared with diagrams produced in other studies and therefore the agreement with the
literature can be analysed. Suggestions for future investigations will also be considered.

4.1 Minkowski Spacetime


The conformal diagrams obtained for the Minkowski spacetime agree exactly with the literature.
This is a frequently explored spacetime and its simplicity means that it is reasonably well under-
stood. The method of using a false radius when reintroducing the angular coordinates dθ and dφ
is a useful tool which aids in the comprehension of three-dimensional plots. This technique is also
utilised by Griffiths and Podolský [2009] when analysing the conformal structure of Minkowski
spacetime.

4.2 De Sitter and Anti-de Sitter Spacetimes


The method used in this report when investigating the geodesics of the de Sitter and anti-de
Sitter spacetimes is different to the method used in the other three main cases (Minkowski,
FLRW and Schwarzschild). Expressions for the Lagrangians are found and then the Euler-
Lagrange equations are used to find geodesic equations. This is a valid method for producing
Penrose diagrams but it is distinct from the methods most commonly found in other works that
study these spacetimes. In most examples, the new timelike and spacelike coordinates are plotted
as straight lines on the diagrams to facilitate simplicity. Other differences arise due to the variety
of potential coordinate systems that can be used. In addition to these variables, this report only
investigates de Sitter spacetime with positive curvature (k = +1).

31
4.3 Robertson-Walker Spacetime
The Penrose-Carter diagrams produced for the three Robertson-Walker spacetimes with angular
coordinates suppressed all agree explicitly with the literature. The introduction of the angular
coordinates to this spacetime has not been found in any other publications. However, the results
obtained here seem to be accurate. As with some of the other spacetimes studied in this report,
for the FLRW universe with k = 0 a false radius is used when reintroducing dθ and dφ, so these
are the only Robertson-Walker Penrose diagrams that do not strictly show a true representation
of the spacetime.
In any future work that may be conducted, the conformal diagrams for FLRW spacetimes
with Λ 6= 0, which are shown in the Background Physics section, could be reproduced. This
could also include the introduction of the angular coordinates to these spacetimes.

4.4 Schwarzschild Spacetime


The Schwarzschild spacetime is both a useful and interesting metric to study, and as a result
it is fairly well covered in the literature. All of the diagrams produced for this spacetime agree
with this literature, including the representation in Kruskal-Szekeres coordinates as well as the
conformal diagrams.
A consideration for future research could be to focus on analysing just the region of this
spacetime on the exterior of the massive body.

4.5 Topological Defects from the Multiverse


The Penrose diagram for a loop defect nucleating in the parent vacuum of a multiverse was
successfully reproduced, based on the research conducted by Zhang et al. [2015]. This was
then expanded upon to produce more conformal diagrams for loops with angles α > 0. This
extension of the analysis is new research and therefore cannot be compared to any existing results.
Nevertheless, the diagrams produced seem to be intelligible and logical.
The probability of a partial bubble-defect collision decreases as the angle of inclination in-
creases. This makes sense, because the loop is effectively shrinking in the x-direction, so it is
much more likely to fully enter the bubble than only partially enter. This is also why the area
of the red region of the de Sitter diagram begins to dominate with increasing angle. At angles of
α > π/4, all of the regions that satisfy the conditions for any collision to occur begin to shrink
until the total area covered becomes 0 for α = π/2. This is also logical, as the loop size in the
x-direction will shrink until it becomes point-like at α = π/2.
An interesting consequence of increasing the angle above zero is the growth of the proportion
of potential collisions on the right of the bubble in comparison to the decrease in the probability
of a collision occurring on the left. This is due to the assumption that the loop has a nucleation
centre on the right of the bubble (rn > 0). The greater the angle of inclination, the lower the
probability that the loop can encircle the bubble and collide on the left, and the higher the
probability that the loop will be restricted to the right side of the bubble and collide on the same
side as its nucleation centre.
In this report, the angled loop was represented by projecting an ellipse onto the x-y plane.
There could be other ways of depicting the loop which could be explored in future research.
Also, it may be possible to investigate a collision when considering the full loop rather than a
projection, or examining collisions in other directions than along the x-axis.

32
5 Conclusion
In this report, the solutions of Einstein’s general theory of relativity have been explored, by
constructing Penrose-Carter diagrams to study the differences between separate spacetimes. The
properties of seven spacetimes have been discussed, and Penrose diagrams have been produced
for five of these solutions.
The mathematical processes used to compactify each spacetime have been fully explained,
describing the steps taken to map infinite space and time on a finite diagram. The use of redefined
timelike and spacelike coordinates has been described, and the relationships between these and
the original coordinates have been expressed.
Various attributes of the spacetimes have been observed in each case, for example the different
curvatures in the Robertson-Walker metric and the horizons in the Schwarzschild spacetime.
The equations that describe these spacetimes include angular coordinates, which have been
reintroduced and analysed in three-dimensional conformal diagrams.
The conformal diagrams produced in this report agree with those found in previous publica-
tions, and new diagrams are created that were not found in any other literature, such as some
of the figures with the angular coordinates dθ and dφ reintroduced.
The research into bubble-defect collisions in the multiverse conducted by Zhang et al. [2015]
has been explained and the results obtained in that paper have been reproduced. The effects
of introducing a loop inclination angle that is greater than zero have been explored, therefore
expanding on previous studies.
Future work could potentially further the results produced in this report by investigating
new solutions, or focussing on specific regions or parameters within the spacetimes explored
here. The research into the topological defects from the multiverse could also be furthered by
attempting to analyse the conditions for a collision between the bubble and a full loop, rather
than a projection of the loop, if this is possible.
Overall, this project has provided some interesting and useful results, including the repro-
duction of calculations and diagrams from existing literature as well as original research and the
discussion of new concepts.

6 References
I. Bengtsson. Anti-de Sitter space. Department of Physics (Stockholm University), 1998.

I. Bengtsson. Conformal compactification and anti-de Sitter space. Department of Physics


(Stockholm University), 2011.

L. Bergström and A. Goobar. Cosmology and Particle Astrophysics. Springer Praxis Books,
2008.

M. Blau. Lecture notes on general relativity. Albert Einstein Center for Fundamental Physics
(Institute for Theoretical Physics, University of Bern), 2014.

S. Carroll. Lecture notes on general relativity. Institute for Theoretical Physics (University of
California), 1997.

J. Frauendiener. Conformal infinity. Living Reviews in Relativity, 7, 2004.

J. Griffiths and J. Podolský. Exact Space-Times in Einstein’s General Relativity. Cambridge


University Press, 2009.

J. Hartle. Gravity: An Introduction to Einstein’s General Relativity. Addison-Wesley, 2002.

S. Hawking and G. Ellis. The Large Scale Structure of Space-Time. Cambridge University Press,
1973.

33
A. Jadczyk. On conformal infinity and compactifications of the Minkowski space. Advances in
Applied Clifford Algebras, 21:721–756, 2010.

V. Mukhanov. Physical Foundations of Cosmology. Cambridge University Press, 2005.

R. Nemiroff and B. Patla. Adventures in Friedmann cosmology: A detailed expansion of the


cosmological Friedmann equations. American Journal of Physics, 76:265–276, 2008.

J. Solà. Cosmological constant and vacuum energy: Old and new ideas. Journal of Physics:
Conference Series, 453, 2013.

M. Spradlin, A. Strominger, and A. Volovich. Les Houches lectures on de Sitter space. Department
of Physics (Harvard University), 2001.

Z. Sun and Y. Tian. Doubled conformal compactification. Science China: Physics, Mechanics
and Astronomy, 57:1630–1636, 2014.

J. Zhang, J. Blanco-Pillado, J. Garriga, and A. Vilenkin. Topological defects from the multiverse.
2015.

34

You might also like