Você está na página 1de 58

Stochastic Gravitational Waves from a new type of

modied Chaplygin Gas


Pedro Ant onio Pinto Frazao
Dissertac ao para a obtenc ao de Grau de Mestre em
Engenharia Fsica Tecnol ogica
J uri
Presidente: Prof. Jo ao Seixas
Orientador: Prof. Alfredo Barbosa Henriques
Co-orientadora: Dra. Mariam Bouhmadi-Lopez
Vogais: Prof. Jose Pedro Mimoso
Novembro 2009
ii
Acknowledgements
This scientic work was done at the Multidisciplinary Center for Astrophysics (CENTRA-IST) in col-
laboration with Dr. Mariam Bouhmadi-Lopez and Prof. Alfredo Barbosa Henriques (Instituto Superior
Tecnico - Universidade Tecnica de Lisboa), to whom I am sincerely grateful.
I would like to thank all the Professors for the interesting things that I have learned during these
years about this great science: Physics.
And, nally, a special mention to my family, and also to all my friends, for the support they gave me
during these years.
iii
iv
Resumo
As ondas gravitacionais de origem cosmologica sao produzidas devido a utuacoes quanticas do vacuo
durante a expansao do universo, em particular nas primeiras fases de forma cao, e, por este motivo,
poderao vir a ser uma fonte importante de informacao acerca do universo primordial.
Neste trabalho investigamos as possveis assinaturas no espectro de energia do fundo estocastico de
ondas gravitacionais durante a transicao do regime inacionario para a fase de domnio de radiacao
do Universo, assumindo uma nova unica cao para o universo primordial. Tal unicacao e no fundo
semelhante `a introduzida por Kamenshchik et al. [1] (ver tambem Refs. [2, 3]), para unicar materia e
energia escuras.
Tal objectivo e alcancado introduzindo um novo modelo cosmologico que corresponde a uma modi-
cacao na equacao de estado, ou, equivalentemente, na densidade de energia, conduzindo a um novo tipo
de gas de Chaplygin generalizado [4]. Neste modelo existe uma transi cao suave entre um universo do tipo
de Sitter, e uma epoca de domnio de radiacao, onde o conte udo de materia e modelado por um udo
exotico, ou, num formalismo equivalente, por um campo escalar.
Desta forma, para realizar um estudo sobre o espectro de ondas gravitacionais, utilizamos o metodo
dos coecientes de Bogoliubov derivado por Parker [5], e usado pela primeira vez por Allen [6], para
calcular o respectivo espectro de energia, tal como seria medido na actualidade, mostrando que, para a
gama de altas frequencias, o espectro depende fortemente de um parametro do nosso modelo.
Por outro lado, usando o n umero de e-folds, o espectro de potencia e ndice espectral das perturbacoes
escalares, podemos efectuar um constrangimento ao modelo cosmologico, usando medicoes da Radiacao
Cosmica de Fundo de Microondas, provenientes do WMAP [7].
Palavras-chave: Chaplygin Gas, Inacao, Ondas Gravitacionais Cosmologicas, e Coecientes
de Bogoliubov.
v
vi
Abstract
The gravitational waves of cosmological origin, due to quantum uctuations of the vacuum, are
produced during the expansion of the universe, in particular from the early phases, and, therefore, they
can give us important information about the very early universe.
In this work, we investigate the possible imprints in the power spectrum of the stochastic background
of the gravitational waves, during the transition from the inationary regime to the radiation dominated
era of the Universe, assuming a new unied early scenario. Such unication is similar in spirit to that
introduced by Kamenshchik et al. [1] (see also Refs. [2, 3]), to unify the dark sectors of the universe, i.e.
the dark matter and the dark energy.
This unication is accomplished through the introduction of a new cosmological model. This model
corresponds to a new type of modied generalised Chaplygin gas [4], because it requires a modication
in the equation of state, or, equivalently, in the energy density. The transition, described by this model,
from an early de Sitter-like stage to a radiation dominated era of the Universe, is very smooth. Finally,
the matter content is modelled by this exotic background uid, with an underlying scalar eld description.
Then, in order to study the gravitational waves, we use the method of the Bogoliubov coecients
introduced by Parker [5], and rst used by Allen [6], to calculate its power spectrum as it would be
measured today. We show that, at the high frequencies range, the spectrum depends strongly on a
parameter of our model. On the other hand, using the number of e-folds, the power spectrum and spectral
index of the scalar perturbations, we constrain this new cosmological model using CMBR measurements
coming from WMAP [7].
Keywords: Chaplygin Gas, Ination, Cosmological Gravitational Waves and Bogoliubov coef-
cients.
vii
viii
Contents
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iii
Resumo . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v
Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii
List of Tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xi
List of Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xiv
I Introduction 1
1 Introduction 3
1.1 Elements of FRW Cosmology and Ination . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 Generalised Chaplygin Gas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.3 Inationary Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.4 Gravitational Waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.5 Outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
II Cosmological Model 9
2 The Model 11
2.1 Energy Density . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.2 Cosmological Evolution - Cosmic Time t . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.3 Cosmological Evolution - Conformal Time . . . . . . . . . . . . . . . . . . . . . . . . . . 13
3 Inationary Dynamics 15
3.1 Inaton Field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3.2 Number of e-folds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
3.3 Slow-roll Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
4 Observational Constraints 19
4.1 Power Spectrum of Density Perturbations . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
4.2 Spectral Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
4.3 Comoving Wavenumber . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
4.4 n
S
r Parameter Space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
III Gravitational Waves 23
5 Gravitational Waves 25
5.1 Power Spectrum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
5.2 Cosmological Evolution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
5.3 Numerical Simulations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
5.3.1 Numerical Integration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
5.3.2 Energy Scale . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
5.3.3 - Parameter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
5.3.4 CMBR/LSS Constraints . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
ix
IV Conclusions 35
6 Conclusions 37
A Numerical Accuracy 39
A.1 Gravitational Energy Spectrum for = 1.03 with CMBR/LSS Constraints . . . . . . . . 40
x
List of Tables
3.1 First range for the -parameter, 1.40 1.10 with steps = 0.05. . . . . . . . . 17
3.2 Second range for the -parameter, 1.23 1.01 with steps = 0.01. . . . . . . . 18
5.1 The respective inationary scale V
0
, in 10
16
GeV, from Eq. (4.3), for the values of and
N
c
of gure (5.7). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
xi
xii
List of Figures
1.1 Current limits (solid lines) and projected sensitivities (dashed lines) to a stochastic gravitational-
wave background versus the gravitational-wave frequency. The dotted curves are scale-
invariant spectra. (From Ref. [27]) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
3.1 Plot of the potential (3.5), as a function of the scalar , for = 1.04 . . . . . . . . . . . 15
3.2 Evolution of and for the range of the -parameter of Table (3.1). The bluer lines
correspond to values that are closer to = 1, and there is a soft black line that represents
the unity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.3 Evolution of and for the range of the -parameter of Table (3.2). The bluer lines
correspond to values that are closer to = 1, and there is a soft black line that represents
the unity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.4 Evolution of (, ), for the range of the -parameter of Table (3.2), in a log-log plot. . . 18
3.5 Evolution of [(, )[ for the range of the -parameter of Table (3.2), in a log-log plot. . . 18
4.1 Constraints on the n
S
r parameter space from CMBR/LSS measurements, for several
models, with the tensor-to-scalar ratio plotted on a log scale, from Ref. [38]. . . . . . . . . 19
4.2 Variation of the inationary scale with the number of e-folds in the range 47 < N
c
< 62.
The limits correspond to N
c
= 47 in blue, and N
c
= 62 in purple. The dashed red line
corresponds to a second method for getting the value
c
through the mode function, see
Section (4.3). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
4.3 Constraints on the n
S
r parameter space for two values of the number of e-folds: N
c
= 47
(blue), and N
c
= 62 (purple). The lines in black are for a variation with a constant -
parameter, parameterized by the number of e-folds, 47 N
c
62. . . . . . . . . . . . . . 22
4.4 The slow-roll parameter evaluated at the moment the mode k
c
exits the horizon, for two
values of the number of e-folds, N
c
= 47 (blue), and N
c
= 62 (purple), and for the method
of using the mode function to calculate
c
(dashed red). . . . . . . . . . . . . . . . . . . 22
5.1 Potential a

/a for all the cosmological evolution (blue), and the Hubble co-moving wavenum-
ber k
2
H
(purple). Dashed gray lines refer to the method of integration: from a starting
inationary phase, a
i
, to a
int
, where it ends the GCG period and begins CDM, until a
f
,
the end of the integration. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
5.2 Evolution [[
2
, proportional to the number of gravitons created, in terms of the scale factor,
and a zoom for the end of integration of large modes. These results are for = 1.03 and
for a energy scale of 1.4 10
16
GeV, given by Eq. (4.3) and using the mode function to
calculate
c
. The second gure stands for a zoom between a
eq1
and a
eq2
. . . . . . . . . . 29
5.3 Spectrum for the energy scales E = 0.110
16
(red), 0.510
16
(orange) and 1.510
16
GeV
(yellow), with = 1.4 . The spectrum suers a upward shift due to an increase in the
energy scale. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
5.4 The potential a

/a for the energy scales E = 1.5 10


16
, 0.5 10
16
and 0.1 10
16
GeV,
in blue, with = 1.4 . The potential suers a reduction with a decrease in the energy
scale. Dashed gray lines refer to the method of integration. It is also shown the variation
in the comoving wavenumber k
2
H
, in purple. . . . . . . . . . . . . . . . . . . . . . . . . . . 30
5.5 Gravitational-wave spectra for the values = 1.03, 1.04, 1.05, 1.10, 1.40, fol-
lowed by = 2.00, and 3.00. The value of moves away from 1 as we move to redder
lines. The spectrum suers an increase in the high frequency region with an increase in
the value of [[. All the curves are calculated for an energy scale of 1.5 10
16
GeV. . . . 31
xiii
5.6 The potential a

/a for = 1.4, 1.3, 1.2, and 1.1 , with the energy scale of 1.0
10
16
GeV. The potential suers a clear decrease in the maximum with a decrease in the
[[-value. It is also shown the variation in the comoving wavenumber k
2
H
, in purple. . . . 31
5.7 Gravitational-wave spectra for the values = 1.05 (red), 1.04 (orange), and
1.03 (yellow) , with N
c
= 47 (solid) and N
c
= 62 (dashed). The spectrum suers a
downward shift as moves away from 1, and, simultaneously, an increase in the high
frequency region. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
5.8 Gravitational-wave spectra for the values = 1.05 (red), 1.04 (orange), and
1.03 (yellow) . The spectrum suers a downward shift as moves away from 1, and, at
the same time, an increase in the high frequency region. . . . . . . . . . . . . . . . . . . . 33
A.1 Gravitational-wave energy spectrum for = 1.03, with an inationary scale given by
CMBR/LSS constraints; and the condition [[
2
[[
2
evaluated for each point of the
spectrum. To each point in the spectrum it corresponds a [[
2
evolution, with the respective
color, in Figure (A.2). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
A.2 Evolution of [[
2
for each mode calculated in the spectrum, from an initial moment when
the mode exits the horizon, until it reenters again. . . . . . . . . . . . . . . . . . . . . . . 40
A.3 Relative error in the derivative of [(a)[
2
, with respect to the scale factor, for each mode
calculated in the spectrum of Fig. (A.1). . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
A.4 Relative error in the derivative of X(a), with respect to the scale factor, for each mode
calculated in the spectrum of Fig. (A.1). . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
xiv
Part I
Introduction
1
Chapter 1
Introduction
1.1 Elements of FRW Cosmology and Ination
Cosmology is in a new era, in which, for the rst time, it is becoming possible to make detailed quantitative
tests of models of the early universe, through cosmological observations. This observational data comes
from experiments such as COBE and WMAP for the Cosmic Microwave Background Radiation (CMBR),
and, in the future, LIGO and LISA, attempting to directly detect the Gravitational Waves (GW). Such
observations are at the present the most plausible route towards the possibility of testing some of the
speculative ideas which, in the recent years, have generated much interest, and also to learn more about
the aspects of physics at extremely high energies.
One of the most important paradigms in the early universe is the so-called inationary era, which
assumes a period of accelerated expansion in the universes distant past. Ination was introduced by A.
Guth [8] and A. Starobinsky [9], as a possible solution to a large number of problems, directly related
with the foundations of the big-bang model, such as the horizon, atness, and the relic density problem.
The inationary paradigm provides an excellent way to solve these problems, and it has also the useful
property that it generates spectra of both density perturbations and gravitational waves. For example,
quantum uctuations during the inationary epoch become macroscopic density uctuations, which leave
distinct imprints in CMBR, and are seeds for Large-Scale Structure (LSS).
The standard cosmology lies in the cosmological principle of a homogeneous and isotropic universe at
large scales. Then, in an appropriate coordinate system, the Friedmann-Robertson-Walker (FRW) metric
becomes
ds
2
= dt
2
+a
2
(t)
_
dr
2
1 Kr
2
+r
2
_
d
2
+ sin
2
d
2
_
_
, (1.1)
where a(t) is the scale factor with respect to the cosmic time t, and the constant K is the spatial
curvature, leading to a closed, at, and hyperbolic spatial geometry, for positive, zero, and negative
values, respectively. Substituting the FRW metric in the well known Einsteins equations
G


1
2
g

R =
8
m
2
P
T

, (1.2)
where m
p
= 1/

G = 1.22 10
19
GeV, in natural units, one gets the Friedmann equations:
H
2
=
8
3m
2
P

K
a
2
, (1.3)
a
a
=
4
3m
2
P
( + 3p) , (1.4)
where a dot denotes the derivative with respect to t, H = a/a is the Hubble parameter. These relations
lead to the equation for the conservation of energy,
+ 3H( +p) = 0 . (1.5)
Dening the density parameter as the ratio of the energy density to the critical density,

c
, with
c

3H
2
m
2
P
8
, (1.6)
3
we can rewrite the Friedmann equation (1.3) in the form
1 =
K
a
2
H
2
. (1.7)
Let us now assume that the equation of state, for the matter in the universe, has the form p
i
= w
i

i
,
where w
i
is a constant, and i refers to a particular type of matter. This includes the main types of matter
that have an important role in cosmology, namely:
radiation, w
r
= 1/3;
non relativistic matter, w
m
0;
cosmological constant, w

= 1.
Then, for a at universe (K = 0), it is easy to combine the Friedmann equation (1.3) with the
energy conservation (1.5), in order to determine the cosmological evolution. For the energy density we
get
i
a
3(1+
i
)
, which, for radiation and non relativistic matter, leads to power-law expansions:
a(t) t
2/3
and a(t) t
1/2
, respectively. In the last case, for the cosmological constant, w

= 1, the
evolution is an exponential one: a(t) e
H t
.
Therefore, one can then rewrite the rst Friedmann equation (1.3) as a sum over the dierent types
of matter in the universe
1
_
H
H
0
_
2
=

i0
_
a
a
0
_
3(1+w
i
)
, (1.8)
which, for the three mentioned types of matter, becomes
_
H
H
0
_
2
=
r0
_
a
a
0
_
4
+
m0
_
a
a
0
_
3
+

. (1.9)
This form will be used in this work to describe the late time evolution of the universe, and corresponds
to the well known CDM model.
The so-called concordance model, or CDM (-Cold Dark Matter), is the simplest model which
adequately ts the observational data coming from CMBR/LSS observations. It also agrees with the
accelerated expansion of the universe in the supernovae observations.
In this model, the acceleration of the Universe is described by a cosmological constant , corresponding
to dark energy, and the cold dark matter (CDM), which stands for a non-relativistic type of dark matter
at the epoch of radiation-matter equality. Finally, to include the inationary era, this model is usually
supplemented by some inationary scenario: a new modied generalised Chaplygin gas [4] in our work.
This model also assumes a nearly scale-invariant spectrum of primordial perturbations and a at universe.
However, on the other hand, the CDM model has the well known cosmological constant problems:
the magnitude problem and the coincidence problem. The rst one points out that the observed value of
the cosmological constant is extremely small to be attributed to the vacuum energy of matter elds; and
the second one the question about why, since there is just a very short period of time in the evolution
of the universe in which the energy density of the cosmological constant is comparable with the energy
density of matter, is this happening just at the present time, such that we can observe it.
1.2 Generalised Chaplygin Gas
As we saw in the last section, there is a large evidence that the Universe, at the present time, is dominated
by a small energy density component, with a negative pressure, the so-called dark energy, that leads to the
observed accelerated expansion, behaving like a cosmological constant. While the most evident candidate
for such component is the vacuum energy, another possible alternative is quintessence [10].
However, instead of a cosmological constant, or the dynamics of a scalar eld rolling down an under-
lying potential, the evidence for a dark energy component can also be explained by a perfect uid with a
modied equation of state [1], avoiding the ne-tuning problems associated with CDM and quintessence
models.
1
The subscript 0 refers to quantities evaluated at the present time.
4
Such modication consists in the introduction of an exotic background uid, the Chaplygin gas,
described by the equation of state [1],
p =
A

, (1.10)
where A is a positive constant, and = 1. Inserting this equation of state into the equation of the energy
conservation (1.5), we get an energy density that scales with the scale factor in the form [1],
=
_
A+
B
a
6
, (1.11)
where B is also a positive constant. Therefore, this energy density interpolates between a dust dominated
phase, where

Ba
3
, at small scale factors, and a de Sitter phase at large scale factors, where p .
Subsequently, this model admits a generalisation which can be regarded as a unication of dark matter
and dark energy.
If we consider the constant a free parameter [1, 2, 3], with values lying in the range 0 < 1, we
get an interpolation between a universe dominated by dust and a De Sitter stage via a phase described
by another equation of state, p = , with ,= 1, behaving like soft matter. Then, from the equation
of the energy conservation (1.5) and the generalised equation of state Eq. (1.10), with in the range
0 < 1, we get the generalised version of Eq. (1.11), the so-called generalised Chaplygin gas (GCG):
=
_
A+
B
a
3(1+)
_ 1
1+
. (1.12)
In the last years, the GCG has received a large attention, and has been constrained with various cosmo-
logical observations [11, 12, 13, 14, 15]. It seems to describe the eective behaviour of dark energy, more
than that of dark matter, and it accounts also the advantage that some GCG models can be excellent
frameworks to analyse dark energy related singularities [17, 18, 19, 20, 21, 16].
In this work we propose a phenomenological model for the inationary epoch and the subsequent
radiation dominated era of the universe, through a unication similar to the one of the Chaplygin gas,
but with respect to the rst periods of the universe.
1.3 Inationary Dynamics
In a large range of inationary models, the implicit dynamics is simply that of a single scalar eld
the inaton rolling down in some underlying potential. In the following, we will see the equations
governing the dynamics of the scalar eld, , whose potential energy, V (), can lead to the accelerated
expansion of the universe.
Neglecting spatial gradients, the energy density, , and the pressure, p, for a homogeneous single
scalar eld are given by
=
1
2

2
+V () , (1.13)
p =
1
2

2
V () . (1.14)
Substituting Eqs. (1.13) and (1.14) into Eqs. (1.3) and (1.5), we obtain,
H
2
=
8
3m
2
P
_
1
2

2
+V ()
_
, (1.15)

+ 3H

+V

() = 0 , (1.16)
where V

dV/d.
These last two equations do not always give an accelerated expansion of the inationary era, but
only in specic ranges, like in the so-called slow-roll regime. The condition for ination, + 3 p < 0,
requires

2
< V () or, classically, that the potential energy dominates over the kinetic energy. Hence,
one requires a suciently at potential for the inaton in order to lead to sucient ination. Then,
5
the slow-roll conditions are given by

2
/2 V () and [

[ 3H[

[, and Eqs. (1.15) and (1.16) are


approximated by
H
2

8
3m
2
P
V () , (1.17)
3H

V

() . (1.18)
The so-called slow-roll parameters are dened by [22]

m
2
P
16
_
V

V
_
2
,
m
2
P
8
V

V
,
2

m
4
P
64
2
V

V
2
. (1.19)
Therefore, the above slow-roll approximations are valid when 1 and [[ 1 for a long period of time.
and also [
2
[ 1, and the inationary phase ends when and [[ approach one.
A very useful quantity, to describe the amount of ination, is the number of e-folds of expansion
before the end of ination, dened by
N ln
a

a
=
_
t

t
Hdt, (1.20)
from a given moment t, and a corresponding scale factor a, to the end of ination at t

with a = a

. Or,
equivalently, in terms of a scalar eld, using Eqs. (1.17) and (1.18), from a given value that increases
until the end of ination at =

:
N
8
m
2
P
_

V
V

d. (1.21)
By denition, the number of e-folds is zero at the end of ination.
The power spectrum, to be evaluated at the moment when a given scale exists the horizon, is deter-
mined by the k-space weighted contribution of modes with given wavenumber. To leading order in the
slow-roll approximation, the amplitudes of the power spectra for scalar perturbations, density perturba-
tions, and tensorial perturbations, i.e. gravitational waves, can be written as [23, 24],
P
s
(k)
128
3m
6
P
V
3
V
2

k=aH
, (1.22)
P
t
(k)
128
3
V
m
4
P

k=aH
, (1.23)
which are a function of the comoving wavenumber k. The scalar eld potential, V , and its derivative, V

,
must be evaluated when a relevant scale in consideration exits the horizon during ination. The power
spectra can also be expanded in power laws,
P
s
(k) P
s
(k
c
)
_
k
k
c
_
1n
s
+(
s
/2) ln(k/k
c
)
, (1.24)
P
t
(k) P
t
(k
c
)
_
k
k
c
_
n
t
+(
t
/2) ln(k/k
c
)
, (1.25)
where k
c
is a pivot wavenumber at which the spectral parameters, the spectral index and its running [22],
n
s
(k) 1 6 + 2 , (1.26)

s
(k) 16 24
2
2 , (1.27)
for the scalar perturbations, and,
n
t
(k) 2 , (1.28)

t
(k) 4 8
2
, (1.29)
for the tensorial perturbations, are to be evaluated. In Section 4.1, we will evaluate P
s
, n
s
, and
s
at the
distance scales of the CMBR and LSS, where they are measured or constrained.
6
To a rst approximation, the power spectra are power laws with power-law indices n
s
and n
t
, although
these indices may run slightly with k, with a running parameterized by
s
and
t
[25]. Finally, using
the slow-roll parameters (1.19) in the expressions (1.22) and (1.23) for the scalar and the tensorial power
spectra, one gets the tensor-to-scalar ratio,
r
P
t
(k)
P
s
(k)
= 16 . (1.30)
1.4 Gravitational Waves
The gravitational waves of cosmological origin, due to quantum uctuations of the vacuum, are produced
during the expansion of the universe, in particular from the early phases, and, therefore, can give us
important information about the very early universe. Despite the fact that the gravitational waves of
cosmological origin have not yet been directly detected, they have received increased attention and there
has been an important eort of research, for their study in the last years.
In this work, we investigate the possible imprints in the power spectrum of the stochastic background
of gravitational waves, for a new model that describes the transition from the inationary to the radiation
dominated eras. This transition is far from being well know, and the signatures associated with it, in
the spectrum, could be a smoking gun leading to the inationary model behind the early stages of
the cosmological evolution. This would give us a new insight into the physics underlying the very early
universe.
Figure 1.1: Current limits (solid lines) and projected sensitivities (dashed lines) to a stochastic
gravitational-wave background versus the gravitational-wave frequency. The dotted curves are scale-
invariant spectra. (From Ref. [27])
Then, we integrate numerically the dierential equations, derived by Parker [5], for the time depen-
dent continuous Bogoliubov coecients, that allow to model all the transitions between distinct periods
of the cosmological evolution as continuous. Hence, this method has advantages over the sudden transi-
tion approximation, because, associated with it, there is always an overproduction of gravitons of large
frequencies, requiring an explicit cut-o for frequencies above the rate of expansion of the universe [6],
which appears in a natural way in the method of the Bogoliubov coecients [26].
7
This last advantage is an important feature because, as we will see, the high frequency range of our
model depends strongly on a parameter of our model, and it is important to have this range without
any numerical artifact. This is also a very useful method to calculate the full spectrum, from the low
frequencies, corresponding to the present cosmological horizon, to the large ones, directly related with
the ination to radiation transition.
Although the present sensitivities of the gravitational-wave detectors, for the relative energy spec-
trum of gravitational waves, do not include the high frequency range, as be can seen in the Fig. 1.1, there
has been a considerable eort to improve these sensitivities like, for instance, in the NASAs Big Bang
Observatory (BBO) [28], the japanese Deci-hertz Interferometer Gravitational Wave Observatory (DE-
CIGO) [29], the Laser Interferometer Gravitational-Wave Observatory (LIGO) [30], and, in the future,
the space-based Laser Interferometer Space Antenna (LISA) [31].
1.5 Outline
This work is organized as follows. After this last review, Part I, of the basic and relevant issues related
with the principal topics for this work, we present a new cosmological model in Part II, and derive the
subsequent cosmological evolution by an analytical integration of the Friedmann equation. Then we
calculate some inationary parameters, from the equivalent scalar eld potential, and we constrain our
model using data coming from CMBR/LSS.
In Part III, we present the method of the Bogoliubov coecients and the cosmological evolution from
the inationary era to the present time, and compare the results of our numerical simulations for the
gravitational wave spectra resulting from dierent combinations of the parameters of our model. Finally,
in Part IV, we summarize and make some concluding remarks.
8
Part II
Cosmological Model
9
Chapter 2
The Model
2.1 Energy Density
In this work, it is proposed a new cosmological model to describe the transition between the inationary
regime and the radiation dominated era of the early universe, leading to a unication of these two stages
of the early evolution, and a transition which evolves in a smooth and instantaneous way. In this model,
the matter content is modelled by a new modied generalised Chaplygin gas, with an underlying scalar
eld description. Such unication is similar to that introduced by Kamenshchik et al. [1], but, in this last
case, to unify the dark sectors of the universe, i.e. the dark matter and the dark energy.
With the above taken into consideration, the new cosmological model, that describes such transition,
corresponds to a simple perfect uid with an energy density evolving as
=
_
A+
B
a
4(1+)
_ 1
1+
, (2.1)
where A and B are constants, necessarily positive, and is a free parameter of the model, that must
satisfy the condition 1 + < 0 in order to obtain an interpolation, between a positive cosmological
constant at small scale factors, and a radiation uid at large scale factors
1
:
A
1
1+

i
, a a

, (2.2)

B
1
1+
a
4

r
, a a

, (2.3)
where
i
is the energy density of the inationary regime,
r
is the energy density during the radiation-
dominated era, and a

is the value of the scale factor at the end of ination, which it is determined by
the condition for the end of ination, i.e. + 3p = 0.
From these two limits of the energy density, we can determine the values of the constants A and B:

i
= A
1
1+
A =
1+
i
, (2.4)

r
=
r0
_
a
0
a
_
4
=
B
1
1+
a
4
B =
_

0
a
4
0
_
1+
. (2.5)
Therefore, the energy density (2.1) can be written in the more intuitive form =
_

1+
i
+
1+
r
_ 1
1+
,
showing the interpolation between the de Sitter and radiation stages.
Then, an early universe, dominated by this uid, would experience a primordial inationary era
through a de Sitter expansion, because it automatically leads to an asymptotic phase where the equation
of state is dominated by a cosmological constant
8
3m
2
P
A
1
1+
. Finally, the universe exits this accelerated
expansion at a = a

, entering in a radiation dominated epoch, with an almost instantaneous transition


between the two stages.
Using this energy density (2.1) in the energy conservation equation (1.5), we can derive the equation
of state satised by its pressure, p, and energy density, ,
p =
1
3

4
3
A

. (2.6)
1
The subscript stands for the value of the quantities at the end of the inationary period.
11
Although this equation of state corresponds to a mixture of a radiation uid and a generalised Chaplygin
gas [2], it is only the total energy density which is conserved. This type of equation of state has been
previously introduced in the context of dark energy models [32, 33, 34, 35], as far as we know initially in
Ref. [32]. However, it has never been analysed in the context of the early-time evolution of the universe.
Notice that, for 0 < < 1, a uid with the energy density (2.1) interpolates between a radiation uid at
small scale factors and a cosmological constant at large scale factor [34].
On the other hand, for < 1, as we saw in the limits (2.2) and (2.3), a uid with the energy density
(2.1) interpolates between a cosmological constant at small scale factors and a radiation uid at large
scale factors. This latter behaviour is the one we are looking for in the present work, as it may correspond
to a viable and a simple phenomenological model for unifying the inationary primordial era and the
radiation epoch of the universe.
2.2 Cosmological Evolution - Cosmic Time t
The evolution of the universe is described by the time variation of the scale factor, that can be determined
using the energy density (2.1) in the Friedmann equation (1.3). Initially, we have a constant Hubble
parameter, since, by the limit (2.2),
H
2

8
3m
2
P
A
1
1+
, (2.7)
and the initial evolution of the universe is described by a de Sitter expansion. The universe expands in
an accelerated way, a > 0, as long as, by Eq. (1.4), we have +3 p 0, where the equality stands for the
end of the inationary period, from which we can determine the corresponding value of the scale factor,
a

=
_
B
A
_ 1
4(1+)
. (2.8)
Thereafter, the scale factor keeps increasing, leading to a radiation dominated universe, with the Fried-
mann equation (1.3) given by
H
2

8
3m
2
P
B
1
1+
a
4
, (2.9)
using the limit (2.3). These two limits have a well known solution that was presented in the Section (1.1),
with power-law solutions.
The general case for the transition (2.1), that we are studying, is more dicult, and there is not an
exact solution. However, we can get an implicit relation between the cosmic time and the scale factor.
More precisely, using the energy density (2.1) in the Friedman equation (1.3) we can do an analytical
integration
2
[36, 37]
2

8
3m
2
P
A
1
2(1+)
(t t

) = (y + 1)
r
F
_
1, r; 1 r;
1
1 +y
_
F (r, r; 1 r; 1) . (2.10)
In the previous expression t is the cosmic time, t

a constant, F(a,b;c;y) corresponds to a hypergeometric


function
3
and
y =
_
a
a

_
4(1+)
, r =
1
2(1 +)
> 0 . (2.11)
The time t t

< 0 measures the cosmic time elapsed since the universe has a radius a < a

until it exits
the inationary era when a = a

at t = t

. Similarly, t t

> 0 measures the cosmic time elapsed since


the universe has a radius a = a

until it reaches a radius a larger than a

. In this model, the universe has


an innite past; in fact the cosmic time reaches innite negative values when a 0 or y 1, because
the hypergeometric function, F
_
1, r; 1 r;
1
1+y
_
, diverges at y = 0. On the other hand, the cosmic
time gets very large values when a 1; i.e. y 1, and the universe becomes radiation dominated.
2
The Friedmann equation (1.3), for introduced in Eq. (2.1), can be integrated in two steps by (i) performing it for
1 + > 0 [36] and (ii) performing an analytical continuation of the hypergeometric function to values of such that
1 + < 0 [37].
3
A hypergeometric series F(b, c; d; x), also called a hypergeometric function, converges at any value x such that |x| 1,
whenever b + c d < 0. However, if 0 b + c d < 1 the series does not converge at x = 1. In addition, if b + c d 1,
the hypergeometric function blows up at |x| = 1 [36].
12
Unfortunately, Eq. (2.10) is quite involved and we cannot obtain an explicit expression of the scale factor
a as a function of the cosmic time t.
By choosing
2

8
3m
2
P
A
1
2(1+)
t

= F (r, r; 1 r; 1) , (2.12)
i.e. by performing a shift on the cosmic time, Eq. (2.10) simplies and reads
(y + 1)
r
F
_
1, r; 1 r;
1
1 +y
_
= 2

8
3m
2
P
A
1
2(1+)
t . (2.13)
Of course, this redenition of the cosmic time does not modify the fact that in this model the universe
has an innite past, where it is asymptotically de Sitter, and an innite future, where it is radiation-
dominated.
2.3 Cosmological Evolution - Conformal Time
As we will see, in order to obtain the gravitational wave power spectrum, it is convenient to introduce
the conformal time (see Eq. (5.15)), , and see if the factor a

/a, where the prime denots a derivative


with respect to conformal time, has a particular simple form. Using the relation dt = a d we can rewrite
Eq. (2.10) in terms of :
a

A
1
2(1+)

8
3m
2
P
(

) = F
_
r,
r
2
; 1
r
2
; 1
_
y

r
2
(1 +y)
r
F
_
r, 1; 1
r
2
;
y
1 +y
_
. (2.14)
At the conformal time

, where the scale factor has the value a = a

, the universe exits the inationary


era. This can be easily checked by noticing that [37]
F
_
r,
r
2
; 1
r
2
; 1
_
= 2
r
F
_
r, 1; 1
r
2
;
1
2
_
. (2.15)
We recall that r, dened in Eq. (2.11), is positive. Therefore, the hypergeometric series in the previous
equality are well dened (see footnote 3). The primordial universe starts its de Sitter expansion at
where a 1. On the other hand, the universe gets radiation-dominated at where
1 a. We note that [37]
F
_
r, 1; 1
r
2
; 1
_
= 1 . (2.16)
Like in the previous section, by choosing
a

A
1
2(1+)

8
3m
2
P

= F
_
r,
r
2
; 1
r
2
; 1
_
, (2.17)
the relation between the conformal time, , and the scale factor given in Eq. (2.14) can be simplied to
a

A
1
2(1+)

8
3m
2
P
= y

r
2
(1 +y)
r
F
_
r, 1; 1
r
2
;
y
1 +y
_
. (2.18)
Again, we notice that this redenition of the conformal time does not modify the fact that in this model
the universe has an innite past (in terms of the conformal time), where it is asymptotically de Sitter,
and an innite future (in terms again of the conformal time), where it is radiation-dominated.
In Section 5.1 it will become clear that, to determine the gravitational wave power spectrum, we have
to solve a set of dierential equations that is equivalent to the second order dierential equation (5.15),
X

+
_
k
2

a
_
X = 0 , (2.19)
in terms of the conformal time. As this equation closely resembles a Shrodinger equation equation with
the potential barrier V a

/a, leading to graviton production. Then, in the following, we will refer to


the factor a

/a as the potential.
For a de Sitter expansion, or a radiation-lled universe, this equation has an exact solution, but, on the
other hand, the intermediate evolution, described by the model (2.1), does not. Hence, the cosmological
evolution of the scale factor in terms of conformal time is implicit given in Eq. (2.18). Therefore, to
calculate the spectrum, it will be necessary a numerical integration of the Eq. (2.19).
13
14
Chapter 3
Inationary Dynamics
3.1 Inaton Field
In order to describe the inationary dynamics of the model presented in the previous section, it is quite
helpful to introduce an equivalent description in terms of a scalar eld, , rolling down the underlying
potential, V (), as we can see in Fig. (3.1). The respective energy density and pressure are given by
(1.13),

=

2
2 a
2
+V () , (3.1)
p

=

2
2 a
2
V () , (3.2)
where the prime stands for derivative with respect to conformal time.
0.0 0.5 1.0 1.5 2.0 2.5 3.0
0.0
0.2
0.4
0.6
0.8
1.0
m
P
V

V
0
Figure 3.1: Plot of the potential (3.5), as a function of the scalar , for = 1.04 .
In a FRW universe, the scalar eld, , is equivalent to the modied Chaplygin gas with the energy
density, , and pressure, p, introduced in Eqs. (2.1) and (2.6), as long as

= and p

= p. Therefore,
the scalar eld, , can be determined analytically in terms of the scale factor, cancelling V () in the
equations (3.1) and (3.2), and, with a change of variable, integrating from a given point to the end of
ination:
(a) =
1

8[1 +[
arcsinh
_
_
B
A
a
2(1+)
_
m
P
. (3.3)
For simplicity, we will restrict to the solution (3.3) with the plus sign (+), as our results do not depend
on which of the two signs we choose.
In a similar way, adding Eq. (3.1) with Eq. (3.2), the scalar eld potential becomes
V () =
1
2
( p) , (3.4)
15
and, using the energy density (2.1) and pressure (2.6), with Eq. (3.3) solved for a = a(), we get the
potential in terms of the scalar eld,
V () =
V
0
3
_
cosh
2
1+
_
8[1 +[ /m
P
_
+ 2 cosh

2
1+
_
8[1 +[ /m
P
_
_
, (3.5)
where V
0
= A
1
1+
.
The universe starts its inationary phase in a de Sitter state where the scalar eld is sitting at the
top of the potential; i.e. at = 0 (see Fig. (3.1)). Then it starts rolling down the potential until it exits
the inationary era when (a

) =

=
1

8[1 +[
ln(1 +

2) m
P
. (3.6)
Finally, the radiation dominated phase starts for large values of the scalar eld, i.e.

.
The derivatives of the scalar eld potential (3.5), with respect to the scalar eld are long expressions
because V () does not have a simple form,
V

=V
0

8
3m
P
cosh
`
8|1 +|/m
P

2
1+
h
1 4 + cosh
`
2

8|1 +| /m
P

i
tanh
`
8|1 +| /m
P

, (3.7)
V

=V
0
16
3m
2
P
cosh
`
8|1 +|/m
P

2
1+

n
cosh
`
2

8|1 +|/m
P

+
h
1 + 4 2(1 + 3) sech
`
8|1 +|/m
P

2
io
, (3.8)
V

=V
0
(8)
3/2
3m
3
P
cosh
`
8|1 +|/m
P

2
1+

h
`
3 + 4
`
1 + 2(1 + 5)

+ 2
`
2 +
2
4
3

cosh
`
2

8|1 +|/m
P

+ cosh
`
4

8|1 +|/m
P

sech
`
8|1 +|/m
P

2
tanh
`
8|1 +|/m
P

. (3.9)
This expressions will be used in Section 3.3 to calculate the slow-roll parameters.
3.2 Number of e-folds
The number of e-folds of expansion, from the moment that a given mode k = aH exits the horizon during
the inationary era, at =
c
, until the end of ination, can be given by (1.21). Using =
c
, we get:
N
c

8
m
2
P
_

c
V
V

d . (3.10)
Substituting the scalar eld potential (3.5) on the previous expression, we obtain
N
c
=
1
2(1 +)(2 1)
ln
8
<
:
"
1 4 + cosh
`
2

8|1 +|
c
/m
P

4(1 )
#
1+
sinh
`
8|1 +|
c
/m
P

3
9
=
;
. (3.11)
Where we can see the dependence of the number of e-folds on the -paramenter of the model, and the
value of the scalar eld when the mode exits the horizon, N
c
= N
c
(,
c
). From this relation, for a given
-value, we have
c
=
c
().
3.3 Slow-roll Parameters
From the denitions of the slow roll parameters (1.19), and using the expressions for the derivatives of the
scalar eld potential derived at the end of Section 3.1, we see that they are functions of the -parameter
of our model and of the scalar eld :
(, ) = tanh
`
8|1 + |/m
P

2
"
cosh
`
8|1 + |/m
P

2
2
cosh
`
2

8|1 + |/m
P

+ 5
#
2
, (3.12)
(, ) =
cosh
`
2

8|1 + |/m
P

+
h
1 + 4 2(1 + 3)sech
`
8|1 + |/m
P

2
i
5 + cosh
`
2

8|1 + |/m
P
, (3.13)
(, ) = tanh
`
8|1 + |/m
P

2
2 sech
`
8|1 + |/m
P

2
1
`
5 + cosh
`
2|1 + |

(

`
3 + 4

h
1 + 2(1 + 5)
i
+ 2
h
2 +
`
1 + + 4
2

i
cosh
`
2

8|1 + |/m
P

cosh
`
4

8|1 + |/m
P

)
. (3.14)
16
The evolution of the rst two slow-roll parameters, and , is plotted in Figures (3.2) and (3.3), for
two dierent ranges of the -parameter, in Tables (3.1) and (3.2), respectively.
This evolution is the rolling-down of the scalar eld, from its starting value, = 0, until it reaches the
end of ination, =

(), which increases with 1 (see Eq. (3.6)). In both Figures the bluer lines
correspond to values that are closer to = 1, and there is a soft horizontal black line that represents
the unity.
The second range is for values much closer to = 1, and, as we can see in these gures, it
corresponds to a better slow-roll approximation, i.e. 1 and 1. The log-log plots of Figures
(3.4) and (3.4) show that, the slow-roll is in fact a better approximation for the second range, and, as it
will be shown in the next section, it is in the second range that the model ts better the observational
constraints.
0.0 0.5 1.0 1.5
0.0
0.2
0.4
0.6
0.8
1.0
1.2
1.4
m
P

0.0 0.5 1.0 1.5


0.0
0.5
1.0
1.5
2.0
m
P

Figure 3.2: Evolution of and for the range of the -parameter of Table (3.1). The bluer lines correspond
to values that are closer to = 1, and there is a soft black line that represents the unity.
-1.40 -1.35 -1.30 -1.25 -1.20 -1.15 -1.10
Table 3.1: First range for the -parameter, 1.40 1.10 with steps = 0.05.
0 5 10 15
0.0
0.2
0.4
0.6
0.8
1.0
1.2
m
P

0 5 10 15
0.0
0.5
1.0
1.5
2.0
m
P

Figure 3.3: Evolution of and for the range of the -parameter of Table (3.2). The bluer lines correspond
to values that are closer to = 1, and there is a soft black line that represents the unity.
17
-1.23 -1.22 -1.21 -1.20 -1.19 -1.18 -1.17 -1.16 -1.15 -1.14 -1.13 -1.12
-1.11 -1.10 -1.09 -1.08 -1.07 -1.06 -1.05 -1.04 -1.03 -1.02 -1.01
Table 3.2: Second range for the -parameter, 1.23 1.01 with steps = 0.01.
0.01 0.05 0.10 0.50 1.00 5.00 10.00
10
6
10
5
10
4
0.001
0.01
0.1
1
m
P

Figure 3.4: Evolution of (, ), for the range of the -parameter of Table (3.2), in a log-log plot.
0.01 0.05 0.10 0.50 1.00 5.00 10.00
10
4
0.001
0.01
0.1
1
m
P

Figure 3.5: Evolution of [(, )[ for the range of the -parameter of Table (3.2), in a log-log plot.
18
Chapter 4
Observational Constraints
In this section, we will contrast our model with the current observational data coming from the cosmic
microwave background (CMBR) and large-scale structure (LSS). It is well known that the measurements
of the inationary observables: the scalar and tensor power-spectrum amplitudes, spectral indices, and
running; that come from CMBR/LSS observational measurements, can lead to constraints on inationary
models (see e.g. [38]).
With measurements from WMAP 5-year data set [39, 40], the Figure (4.1) shows that the large-eld
models, V () =
4
and V () = m
2

2
, are not consistent with the inationary constraints from the
data. On the other hand, there is a large number of models whose parameters, spectral index n
S
and
scalar-to-tensor ratio r, t the observational data, in particular, the small-eld models: natural ination
from a pseudo-Nambu-Goldstone boson, with potential V () =
4
_
1cos (/)

; a logarithmic potential
V () ln (), typical of supersymmetric models; and a Coleman-Weinberg potential V ()
4
ln ().
Figure 4.1: Constraints on the n
S
r parameter space from CMBR/LSS measurements, for several
models, with the tensor-to-scalar ratio plotted on a log scale, from Ref. [38].
The precise constraints to the inationary observables depend strongly on the combination of ob-
servational data sets. In the following, we simply take P
s
(k
c
) = (2.45 0.23) 10
9
, n
s
= 1.0 0.1,
and [
s
[ < 0.04 at a pivot wavenumber k
c
= 0.05 Mpc
1
[41]. It is easy to see that the condition for
[n
S
1.0[ < 0.1 is included in the condition for its running [
S
[ < 0.04, for this mode in particular. Then,
for the contraints on the model, we have to consider only the one on n
S
.
19
4.1 Power Spectrum of Density Perturbations
To rst order in the slow-roll approximation, the amplitude of the power spectrum for density perturba-
tions (1.22), evaluated at the pivot wave number k
c
, is given by
P
S
(k
c
) =
128
3m
6
P
V (
c
)
3
V

(
c
)
2
. (4.1)
This expression determines the value of the prefactor V
0
, which corresponds to the inationary scale. If
we dene V () = V
0
f(, ), where f is a function of the scalar eld that depends on the -parameter
of the model, we obtain
V
0
=
3m
6
P
128
P
S
(k
c
)
f

(,
c
)
2
f(,
c
)
3
. (4.2)
From the corresponding function f(, ), the inationary scale V
0
becomes,
V
0
= 18 P
S
(k
c
)
tanh
_
8[1 +[
c
/m
P
_
2
_
cosh
_
8[1 +[
c
/m
P
_
2
2
_
2
_
cosh
_
8[1 +[
c
/m
P
_
_

2
1+
_
cosh
_
2

8[1 +[
c
/m
P
_
+ 5
_
3
m
4
P
, (4.3)
where we clearly see that V
0
= V
0
(,
c
).
In Section 3.2 we saw that the number of e-folds is a function of and the parameter: N
c
= N
c
(, ).
Then, for a given number of e-folds, N
c
, we have an implicit relation of the form
c
=
c
(), in Eq. (3.11).
Therefore, for a given number of e-folds, and by xing also the -parameter, we have the value of the
scalar eld at the moment the mode exits the horizon,
c
, and we can evaluate, for instance, the scale of
energy through Eq. (4.3): V
0
= V
0
_
,
c
()
_
.
As we will see in Section 4.3, we can use the mode function (4.10), k
H
= k
H
(, ), instead of the
number of e-folds, to determine the value of
c
. For a given -parameter we have
c
from k
c
= k
H
(,
c
),
and then a relation V
0
= V
0
_
,
c
()
_
. The curves that result from this two dierent methods are
presented in Fig. (4.2).
1.05 1.04 1.03 1.02 1.01 1.00
0.2
0.5
1.0
2.0
5.0
10.0

V
0

k
c

1
0
1
6
G
e
V

Figure 4.2: Variation of the inationary scale with the number of e-folds in the range 47 < N
c
< 62.
The limits correspond to N
c
= 47 in blue, and N
c
= 62 in purple. The dashed red line corresponds to a
second method for getting the value
c
through the mode function, see Section (4.3).
20
4.2 Spectral Parameters
From the derived expressions for the slow-roll parameters (3.12), (3.13) and 3.14), we can see that the
spectral index (1.26) and its running (1.27) are functions of the scalar eld and depend on the -parameter:
n
S
=3
288(1 +)
2

5 + cosh

8|1 +|/m
P

2
+
16(1 +)(5 + 2)
5 + cosh

8|1 +|/m
P
4 sech
2

8|1 +|/m
P

, (4.4)

S
= 8(1 +)
tanh
2

8|1 +|/m
P

5 + cosh

8|1 +|/m
P

(
92 + 918 + 296
2
+

106 +(797 + 388)

cosh

8|1 +|/m
P

+
+ 2

2 +(61 + 44)

cosh

8|1 +|/m
P

(2 +)(5 + 4) cosh

8|1 +|/m
P

2 sech
2

8|1 +|/m
P

. (4.5)
4.3 Comoving Wavenumber
There is one more relation that we must consider for this analysis, the one dening the comoving wave
number,
k
H
=
2

H
= 2 a H =

32
3
3m
2
P
a
1/2

1/2
(4.6)
where
H
= (a H)
1
. The energy density can be rewritten as
= A
1
1+
_
1 +
B
A
a
4(1+)
_
. (4.7)
From the scalar eld evolution in terms of the scale factor (see Eq. (3.3)), using the relations from the
expressions (2.4) and (2.5), we obtain
a() =
_

r0

i
_
1/4
a
0
sinh

1
2(1+)
_

8[1 +[/m
P
_
, (4.8)
where the subscript 0 stands for the values of the quantity at the present time. The energy density
reads then
() =
i
cosh
2
1+
_

8[1 +[/m
P
_
. (4.9)
Using Eq. (4.8) and Eq. (4.9) in the coming wavenumber (4.6) we get,
k
H
=

32
3
3m
2
P

1/4
r0

1/4
i
a
0
_
cosh
_

8[1 +[/m
P
_
coth
_

8[1 +[/m
P
__ 1
2(1+)
. (4.10)
Because
r0
is a constant, to conclude that k
H
= k
H
(, ) we must see the dependence of
i
. From the
relation (2.4) we have
i
= A
1
1+
, and, on the other hand, V
0
= A
1
1+
, see Section (3.1). Then
i
V
0
,
and, through the dependence V
0
= V
0
(, ) in Eq. (4.3), we conclude that
i
=
i
(, ), and, nally,
that k
H
= k
H
(, ). From this implicit relation, if we choose the mode k
c
, and a -value in the range
1 + < 0, we have a value for
c
through the relation k
H
(,
c
) = k
c
. This is another method, instead
of the number of e-folds, as already mentioned, to determine
c
in order to evaluate the inationary
observables.
4.4 n
S
r Parameter Space
In Fig. (4.3) we present the n
S
r parameter space for our model, with the spectral parameter, n
S
, and
the tensor-to-scalar ratio, r, given by the expressions in the equations (1.26) and (1.30), respectively.
21
In the same gure we can see that the spectral index is at most of the order of n
S
0.9, for in the
range 1.03 < < 1.02. Comparing with the observational constraints from CMBR/LSS, that result
in bands of dierent condence levels, plotted in Fig. (4.1), 95% in green and 68% in blue, we see that
the bands for our model are far from these ones. We see that the spectrum is slightly more red than
preferred by the observation. However, Fig. 4.3 shows that r < 1 is in agreement with the observation
[24], and our best t value corresponds to = 1.024; i.e. an energy scale for ination about 10
16
GeV.
= -1.03
= -1.04
= -1.02
0.80 0.85 0.90 0.95 1.00
0.0001
0.001
0.01
0.1
1
0.5
1.0
1.5
2.0
3.0
5.0
n
S
r
V
0
1

1
0
1
6

G
e
V

Figure 4.3: Constraints on the n


S
r parameter space for two values of the number of e-folds: N
c
= 47
(blue), and N
c
= 62 (purple). The lines in black are for a variation with a constant -parameter,
parameterized by the number of e-folds, 47 N
c
62.
In fact, the expressions used for n
S
and r are only valid in the slow-roll regime. To see if these
expression are valid when we evaluate these parameters at the moment the mode exits the horizon,
we must have (,
c
) 1. In Figure (4.4) we see that, in the range 1.04 < < 1.02, the slow-
roll parameter obeys that condition. Then, the results in Fig. (4.3) are calculated using the correct
expressions for n
S
and
S
.
1.05 1.04 1.03 1.02 1.01 1.00
10
6
10
4
0.01
1

Figure 4.4: The slow-roll parameter evaluated at the moment the mode k
c
exits the horizon, for two
values of the number of e-folds, N
c
= 47 (blue), and N
c
= 62 (purple), and for the method of using the
mode function to calculate
c
(dashed red).
22
Part III
Gravitational Waves
23
Chapter 5
Gravitational Waves
5.1 Power Spectrum
The tensorial part of the linear perturbation for the metric of a at universe, in conformal time, is given
by
ds
2
= a
2
()
_
d
2
+ [
ij
+h
ij
(, x)] dx
i
dx
j
_
, (5.1)
where h
ij
are the tensorial perturbations, with the latin indices i, j running from 1 to 3, that can be
expanded, in the usual way, in terms of plane waves as
h
ij
(, x) =

8
m
2
P
2

p=1
_
d
3
k
(2)
3/2
a()

2k
_
a
p
(, k)
ij
(k, p)e
ikx
(, k) + H.c.

, (5.2)
where k = [k[ = a is the comoving wavenumber, the sum is over two polarizations states p, of the
gravitational waves, a
p
is the annihilation operator,
ij
the polarization tensor, H.c. stands for the
Hermitic conjugate of the rst term, and the mode function obeys the equation of a parametric oscillator

+
_
k
2

a
_
= 0 , (5.3)
also in conformal time.
The stochastic gravitational wave energy spectrum produced during the expansion of the universe
is calculated by the method of continuous time dependent Bogoliubov coecients. This method uses
the fact that, in the evolution of the universe, the annihilation a
p
(, k) and creation a

p
(, k) operators
change with time. Through a Bogoliubov transformation, these can be written in terms of time-xed
annihilation A
p
(k) and creation A

p
(k) operators
a
p
(, k) = (, k)A
p
(k) +

(, k)A

p
(k) . (5.4)
The two Bogoliubov coecients and must satisfy the normalization condition [[
2
[[
2
= 1.
Using the last equation, and the analogous one to a

p
(, k), it can be shown that the coecient
gives de number of gravitons created: [[
2
= N
k
()) [42]. Using the denitions of the density of states,

2
d/(2
2
c
3
) and of the energy density in terms of the power spectrum, dE = P() d, and considering
the two polarization states we have
P() =

3

2
c
3
[(
0
)[
2
. (5.5)
The relative logarithmic energy spectrum of the gravitational waves, is dened by
(,
p
)
1

c
d
gw
d ln
, (5.6)
where
c
is the value of the present time critical energy density and
gw
is the gravitational wave energy
density

gw
=
_
P()d. (5.7)
25
Hence, the dimensionless relative logarithmic energy spectrum of the gravitational waves (5.6), at the
present time
0
, is proportional to
2
, at the same time:

GW
(,
0
) =
8 m
2
P
3c
5
H
2
(
0
)

2
(
0
) . (5.8)
Therefore, the evolution of the Bogoliubov coecient , from early times to the present time, can give
us the power spectrum at the present time. This can be done by solving the set of coupled dierential
equations [43],

=
i
2k
_
+e
2ik(
i
)
_
a

a
, (5.9)

=
i
2k
_
+e
2ik(
i
)
_
a

a
, (5.10)
Using new functions X(, k) and Y (, k), through the relations
=
1
2
(X +Y )e
ik(
0
)
, (5.11)
=
1
2
(X Y )e
ik(
0
)
, (5.12)
where
0
is an arbitrary constant, that can be put equal to
i
, corresponding to an inationary de Sitter
universe, with a constant Hubble parameter, the equations (5.9) and (5.10) become
X

= ikY , (5.13)
Y

=
i
k
_
k
2

a
_
X , (5.14)
These, in turn, are equivalent to the second order dierential equation
X

+
_
k
2

a
_
X = 0 , (5.15)
with
Y =
i
k
X

. (5.16)
The Friedmann equation shows that, either in cosmic time or conformal time, a =
4 G
3
a ( + 3 p)
and a

=
4G
3
a
3
( 3 p), respectively, the acceleration of the scale factor depends on the matter content
of the universe and its equation of state. Therefore, because the power spectrum at the present time is
calculated through the equation (5.15), that depends on a

/a, it will reect the evolution of the matter


content dominating the universe at each moment. For example, the radiation uid does not increase the
number of gravitons created during the expansion because p = /3 and a

/a = 0, for this case, leading


to an adiabatic variation of the amplitude of the gravitational wave.
After introducing the method used to calculate the gravitational wave energy spectrum, with the equa-
tions to determine the continuous evolving Bogoliubov coecients we, must derive now the appropriate
initial conditions.
As we have seen in Section 2.3, for our model (2.1), that describes the transition between the ina-
tionary period and the radiation-dominated era, the factor a

/a, has not an exact analytical solution,


and we must use numerical methods to solve the set of equations. The integration must start at the
beginning of the inationary period, where it admits an exact analytical solution for the case of a de
Sitter universe. In this case, the set of dierential equations yields a solution that, at the inationary
period, takes the form [42]
X(
i
) =
_
1 +i
a(
i
)H
k
_
e
i
k
a(
i
)H
, (5.17)
Y (
i
) =
_
1 +i
a(
i
)H
k

a
2
(
i
)H
2
k
2
_
e
i
k
a(
i
)H
. (5.18)
where a(
i
) is the scale factor and H(
i
) = H the constant Hubble parameter, at some moment
i
, during
the inationary phase. These equations will be used as initial conditions for X() and Y ().
26
5.2 Cosmological Evolution
In order to calculate the gravitational wave energy spectrum, we must integrate the set of dierential
equations (5.13) and (5.14) from the beginning of the inationary period to the present time, with the
initial conditions given by the solutions (5.17) and (5.18) for a de Sitter Universe. The value of the
Hubble parameter, during ination is directly given by the energy scale E of ination:
H(
i
)
2
=
8
3m
2
P

i
, (
i
)
1/4
= E , (5.19)
in Planck units, and, initially, we take its standard value of E = 1.5 10
16
GeV.
The integration has two stages, a rst one for the transition ination-radiation described by our new
modied GCG [4], and a second one for late time, described by CDM model. From the rst Friedmann
equation (1.3), written in terms of the conformal time, we get
a

a
=
4
3 m
2
P
a
2
_
3p
_
, (5.20)
_
a

a
_
2
=
8
3 m
2
P
a
2
. (5.21)
This set of equations takes a particular form in each one of the integration parts. In the rst part, using
the energy density (2.1) and pressure (2.6), the transition from ination to radiation era is described by
the set of equations
a

a
=
4
3m
2
P
a
2
A
_
A+
B
a
4(1+)
_


1+
, (5.22)
_
a

a
_
2
=
8
3m
2
P
a
2
_
A+
B
a
4(1+)
_ 1
1+
, (5.23)
where the constants A and B are determined by the limits (2.4) and (2.5), respectively.
For this calculation, as we saw, it is important to ensure the right behaviour of a

/a, at the moment


of interchange, a
int
, between the two periods of the cosmological evolution; we must ensure that
_
a

a
_
GCG

_
a

a
_
CDM
. (5.24)
After this, the evolution is determined by the CDM model
a

a
=
4
3m
2
P
a
2
_

m0
_
a
0
a
_
3
+ 4

_
, (5.25)
_
a

a
_
2
=
8
3m
2
P
a
2
_

r0
_
a
0
a
_
4
+
m0
_
a
0
a
_
3
+

_
. (5.26)
The values for the radiation and matter energy densities at the present time are given through the usual
denition of these parameters in terms of dimensionless density parameter , =
3 H
2
0
8
, where
H
0
= 71 km/s/Mpc. The values for these parameters, given by WMAP 5-year [7], are:

r
= 8 10
5

cdm
= 0.223 0.015

b
= 0.0451 0.0017 (5.27)
and

m
= 0.268
+0.016
0.017

= 0.732
+0.017
0.016
(5.28)
We dene the scale factor at the present time to be a
0
= 10
58
, with a = 1 at the beginning of our
integration.
27
Finally, it is important to make a reference to the moments of transition between the radiation
dominated and matter eras, a
eq1
, and between matter era and dark energy dominated era, a
eq2
, that will
be used in the following discussion, and which are given by
a
eq1
= a
0

r0

m0
, (5.29)
a
eq2
= a
0
_

m0

_
1/3
. (5.30)
5.3 Numerical Simulations
In this section, we calculate the full gravitational spectrum of the new modied generalised Chaplygin
gas, using CDM for late times. There are two important parameters to be considered in our model
(2.1): the scale of ination, that xes the value of the positive constant A, and the parameter , in the
range 1 + < 0.
The better way to perform the integration, reminding that the equations for the cosmological evolution
(5.22) and (5.25) are functions of the scale factor, is by taking the scale factor as the independent variable,
d
d
=
d a
d
d
d a
. Then, the system of coupled dierential equations (5.13) and (5.14) can be rewritten in the
form
d X
d a
= i
k
a

Y, (5.31)
d Y
d a
= i
1
k a

_
k
2

a
_
X , (5.32)
where the prime denotes derivative with respect to the conformal time.
Using the scale factor as the indepent variable we avoid the use of the Friedmann equations, Eqs.
(1.3) and (1.4), to determine the evolution of the scale factor with the time. Hence, the time evolution
of the scale factor is implicitly included in the time derivatives of the scale factor, given by Eqs. (5.23)
and (5.26), of the previous set of equations. This time evolution of the scale factor is determined by the
evolution of the matter content of the universe, in particular in the energy density and the pressure.
a
int
a
f
a
i
k
2
Modified Chaplygin Gas CDM
1
10
10
10
20
10
30
10
40
10
50
10
60
10
8
100
10
12
10
22
10
32
10
42
a
a''
a
Figure 5.1: Potential a

/a for all the cosmological evolution (blue), and the Hubble co-moving wavenum-
ber k
2
H
(purple). Dashed gray lines refer to the method of integration: from a starting inationary phase,
a
i
, to a
int
, where it ends the GCG period and begins CDM, until a
f
, the end of the integration.
5.3.1 Numerical Integration
In Fig. (5.1) we present the method of integration. For a given mode k, the integration is done from
an initial scale factor, a
i
, at the GCG inationay period, just before the mode exits the horizon; and
28
continue up to CDM late times, until a moment just after the mode reenters the horizon, corresponding
to a nal scale factor, a
f
.
Therefore, there are two stages of integration, one for the GCG period and another for the CDM
period. In order to preserve the continuity of the factor a

/a in the set of dierential equations, the


interchange between the two stages at a
int
is determined by (5.24).
a
a
eq1
a
eq2
100
10
12
10
22
10
32
10
42
10
52
10
20
100
10
24
10
46
10
68
10
90
a

2
a
eq1 a
eq2
10
50
10
52
10
54
10
56
10
58
10
90
10
93
10
96
10
99
10
102
10
105
a

2
Figure 5.2: Evolution [[
2
, proportional to the number of gravitons created, in terms of the scale factor,
and a zoom for the end of integration of large modes. These results are for = 1.03 and for a energy
scale of 1.4 10
16
GeV, given by Eq. (4.3) and using the mode function to calculate
c
. The second
gure stands for a zoom between a
eq1
and a
eq2
.
The Fig. (5.2) shows the time evolution of [[
2
, in terms of the scale factor, for this method of
integration. Each color corresponds to the integration for a given mode k
2
, from a
i
till a
f
, and the red
lines are for the ones with a small value, or small frequencies. The zoom of the log-log plot, in the same
Fig. (5.2), that referes to dark matter stage between a
eq1
(5.29) and a
eq2
(5.30), shows that there is a
strong contribution of dark matter for the graviton production, that gradually disapears for large modes
(bluer lines in the zoom).
Hence, Fig. (5.2) shows that, the method of taking the end of integration just after the mode reenters
the horizon, cannot be applied for the rst lower frequencies, because there is a graviton production
from the present acceleration of the universe, during the CDM era, which becomes negligible for large
k
2
modes, and, just in this case, the method of integration can be used and also the extrapolation
(a
f
) (a
0
). This approximation, (a
f
) (a
0
), means that we consider the value of the Bogoliubov
coecient approximatelly constant between the end of integration, at a
f
, and the nal value at the
present time a
0
, with a
0
> a
f
. It is this value that we are looking for, in order to determine the
gravitational spectrum at the present time.
On the other hand, taking the initial integration scale factor a
i
as the moment just before the mode
exits the horizon can also be applied as long as this value is far from the transition ination-radiation,
which is almost always true because this transition is approximately instantaneous, as we will see when
studying the continuity of the energy density during all the evolution. Hence, the initial conditions are
X(a
i
) =
_
1 +i
a
i
H
k
_
e
i
k
a
i
H
, (5.33)
Y (a
i
) =
_
1 +i
a
i
H
k

a
2
i
H
2
k
2
_
e
i
k
a
i
H
. (5.34)
Therefore, we conclude that the application of this method is valid, and constitutes a very good way
to save integration time.
The numerical method used for all the integrations was the BDF (Backward Dierentiation Formulas)
of Mathematica, well suited for sti systems like in this problem. The condition [[
2
[[
2
is used as a
constraint to ensure the accuracy of the numerical results, and that any possible signicant imprint, that
may result in the power spectrum, is not a numerical artifact. The relative error for [[
2
and X() and
Y () constitute a further check of this accuracy, which we study in Appendix A.
Finally, we calculate the evolution of (a) and its value at the end of the integration, which, with the
extrapolation to the present time, and, from relation (5.8), give the power spectrum
GW
(a
0
), at the
present time, for each mode k.
The two free parameters of the model are the positive constant A, determined by the energy scale, and
the -parameter, with the condition < 1. In the following sections, we present our results, coming from
29
numerical simulations, for some gravitational waves spectra. These spectra are obtained using dierent
ways of xing the values of these two free parameters: the energy scale and the -parameter. We also
study the subsequent implications associated with the energy scale in Sec. 5.3.2; then the variations in
the spectrum due to the -parameter, Sec. 5.3.3; and, nally, those results consistent with the CMB
constraints, Sec. 5.3.4.
5.3.2 Energy Scale
Some gravitational wave spectra are shown in Fig. (5.3) for dierent values of the energy scale. Here we
can see that, with a decrease in the energy scale, the spectrum suers a downward shift and, at the same
time, a decrease of the high frequency region, which can be associated with the observed variations in
the potential
a

a
, as seen in Fig. (5.4). The decrease in the maximum of the potential, in the ination-
radiation transition, is directly related with the decrease of the high frequency region of the spectrum.
This fact will become clear in the next section.
10
17
10
12
10
7
0.01 1000
10
8
10
23
10
21
10
19
10
17
10
15
f Hz

G
W
Figure 5.3: Spectrum for the energy scales E = 0.1 10
16
(red), 0.5 10
16
(orange) and 1.5 10
16
GeV
(yellow), with = 1.4 . The spectrum suers a upward shift due to an increase in the energy scale.
a
int
a
f
a
i
E
1
10
10
10
20
10
30
10
40
10
50
10
60
10
8
100
10
12
10
22
10
32
10
42
a
a''
a
Figure 5.4: The potential a

/a for the energy scales E = 1.5 10


16
, 0.5 10
16
and 0.1 10
16
GeV, in
blue, with = 1.4 . The potential suers a reduction with a decrease in the energy scale. Dashed gray
lines refer to the method of integration. It is also shown the variation in the comoving wavenumber k
2
H
,
in purple.
30
5.3.3 - Parameter
Besides the implications of the energy scale, another more important issue is the dependence on the
parameter of the model, which is clearly shown in Fig. (5.5). We see that, for a given energy scale,
the high frequency region increases with the [[-value, and, in particular, it increases when the -value
moves away from = 1. There is not a upward shift because the energy scale remains constant. It is
also clear the relation of this dependence with the subsequent variation in the potential, plotted in Fig.
(5.6), where the maximum of the potential decreases for the same variation of the parameter .
10
18
10
13
10
8
0.001 100
10
7
10
19
10
18
10
17
10
16
10
15
10
14
f Hz

G
W
Figure 5.5: Gravitational-wave spectra for the values = 1.03, 1.04, 1.05, 1.10, 1.40, followed
by = 2.00, and 3.00. The value of moves away from 1 as we move to redder lines. The spectrum
suers an increase in the high frequency region with an increase in the value of [[. All the curves are
calculated for an energy scale of 1.5 10
16
GeV.
1
a
int a
f
k
2
1
10
10
10
20
10
30
10
40
10
50
10
60
10
8
100
10
12
10
22
10
32
10
42
a
a''
a
Figure 5.6: The potential a

/a for = 1.4, 1.3, 1.2, and 1.1 , with the energy scale of 1.0
10
16
GeV. The potential suers a clear decrease in the maximum with a decrease in the [[-value. It is
also shown the variation in the comoving wavenumber k
2
H
, in purple.
31
5.3.4 CMBR/LSS Constraints
The nal question concerns the discussion about the CMBR/LSS contraints in Chapter 4. The inationary
scale V
0

i
, given by Eq. (4.3), is used to determine the value of the constant A. But, in order to
determine its value, for a given -parameter, it is necessary to calculate the magnitude of the scalar eld
when the mode k
c
exits the horizon, =
c
. As we saw, there are two implicit relations for
c
=
c
():
a rst one from the number of e-folds, N
c
= N
c
(,
c
), in Eq. (3.11), and a second one through the mode
function (4.10), k
c
= k
H
(,
c
). Then, for a given -value and the respective value of V
0
, we get the
constant A, and the gravitational-wave spectrum can be calculated.
N
c
-1.06 -1.05 -1.04
47 0.423 0.706 1.161
62 0.168 0.325 0.625
Table 5.1: The respective inationary scale V
0
, in 10
16
GeV, from Eq. (4.3), for the values of and N
c
of gure (5.7).
Gravitational-wave energy spectra, for the rst method, are shown in gure (5.7), for dierent values
of , and two numbers of e-folds. The inationary scale used for each line is shown in Table (5.1), is
calculated through Eq. (4.3), for a given number of e-folds, see Eq. (3.11).
To understand this variation it is necessary to consider the inationary scale behaviour as a function
of , plotted in Fig. (4.2). The same picture, shows that an increase in the number of e-folds, from
N
c
= 47 to N
c
= 62, leads to a decrease in the inationary scale V
0
. Hence, from the conclusions of
Section 5.3.2, the decrease in Fig. (5.7), between solid lines, N
c
= 47, and dashed lines, N
c
= 62, is easily
interpreted as the mentioned decrease in the inationary scale.
But, on the other hand, from the discussion of Section 5.3.3, when the -value moves away from 1,
there is an increase in the high frequency region. The same behaviour happens in this case, as we can
see in Fig. (5.7) where, for a given number of e-folds, for example N
c
= 47 (solid lines), an increase in
[[, from = 1.04 to = 1.06, leads to an increase in the high frequency region.
Therefore, the gravitational-wave energy spectra consistent with CMBR/LSS constraints, in Fig.
(5.7), shows a behaviour that is a combination of the conclusions from the last two sections: a downward
shift as a result from the decrease in the inationary scale; and an increase in high frequency region that
results from a increase in the [[-value.
10
18
10
14
10
10
10
6
0.01 100
10
19
10
18
10
17
10
16
10
15
10
14
f Hz

G
W
Figure 5.7: Gravitational-wave spectra for the values = 1.05 (red), 1.04 (orange), and
1.03 (yellow) , with N
c
= 47 (solid) and N
c
= 62 (dashed). The spectrum suers a downward shift
as moves away from 1, and, simultaneously, an increase in the high frequency region.
If, instead of the number of e-folds N
c
, we use the mode function (4.10) to calculate the value of
c
for a given , in order to determine the inationary scale, we get similiar results to the ones from the
other method, wich are shown in Fig. (5.8) for the values = 1.02, 1.03 , and 1.04.
32
To understand this result it is necessary to see also the inationary scale behaviour as a function of
, plotted in Fig. (4.2). This gure shows that the dashed red line for the mode function has a similiar
variation as the solid lines for the number of e-folds. Then, it is clear that the variation in Fig. (5.8)
is also a combination of these results: a variation in the energy scale, because the spectrum suers a
downward shift in consequence of the decrease in the inationary scale; and a variation of alpha, because,
as we move away from = 1, there is an increase in the high frequency region of the spectrum.
10
18
10
14
10
10
10
6
0.01 100
10
19
10
18
10
17
10
16
10
15
10
14
f Hz

G
W
Figure 5.8: Gravitational-wave spectra for the values = 1.05 (red), 1.04 (orange), and
1.03 (yellow) . The spectrum suers a downward shift as moves away from 1, and, at the same
time, an increase in the high frequency region.
33
34
Part IV
Conclusions
35
Chapter 6
Conclusions
In this work, we have investigated the production of gravitational waves in a new modied generalised
Chaplygin gas. This new cosmological model describes a smooth and instantaneous unied transition
between the inationary regime and radiation dominated era. The -parameter of our model was con-
strained using recent measurements from the cosmic microwave background (CMBR) and large-scale
structure (LSS).
Besides the fact that the model is a bit tight with the observational constraints in the n
S
r parameter
space, the obtained spectra reveal a consistent picture corresponding to this smooth transition, and
constitute a signicant imprint of this modied generalised Chaplygin gas. One of the merits of this
transition is its relative simplicity.
We have used the method of continuous Bogoliubov coecients to calculate the gravitational-wave
energy spectrum for dierent scales of energy and values of the -parameter of the model.
Finally, the strong variation of the high frequency range, is directly related with the decrease in the
maximum of the potential, and implies strong limits to the maximum frequency one will be able to observe
in future gravitational-waves detectors, like BBO and DECIGO [27], for the KHz range of frequencies.
In fact, for the more consistent values of -parameter, the spectrum shows a frequency in the Hz region.
In conclusion, these results show that, for this model, the Hz-kHz frequency range comes directly
from the transition between the inationary regime and the radiation era, and addresses important issues
about the limits of the gravitational wave energy spectrum.
37
38
Appendix A
Numerical Accuracy
As already mentioned, the numerical method used for all the integrations was the BDF (Backward
Dierentiation Formulas) of Mathematica, well suited for sti systems like the one in this problem. The
condition [[
2
[[
2
is used as a constraint to ensure the accuracy of the numerical results, and that
any possible signicant imprint, that may result in the power spectrum, is not a numerical artifact. The
relative error for [[
2
and X() and Y () constitute a further check of this accuracy.
In the next section we only present the results for = 1.03, which is the -value that best ts
the observational data, as we saw in Section 4.4. The gravitational wave spectrum, calculated for this
-value and with an inationary scale given by the CMBR/LSS constraints, is presented in Fig. (A.1),
where there are shown the points calculated for the dierent modes k
2

2
.
To ensure that our results, coming from the numerical simulations, are consistent and not a numerical
artifact, we seek the behaviour of the condition [[
2
[[
2
, in Fig. (A.2), and the relative error for [[
2
and X(), presented in Figs. (A.3) and (A.4), respectively. As we can see in these gures, the numerical
constraints are very well satised, and, therefore, the numerical results have indeed a great numerical
accuracy.
39
A.1 Gravitational Energy Spectrum for = 1.03 with CMBR/LSS
Constraints
10
18
10
14
10
10
10
6
0.01 100
10
18
10
17
10
16
10
15
10
14
f Hz

G
W
a
eq1
a
int a
eq2
1
10
10
10
20
10
30
10
40
10
50
10
60
0.4
0.6
0.8
1.0
1.2
1.4
a

2
Figure A.1: Gravitational-wave energy spectrum for = 1.03, with an inationary scale given by
CMBR/LSS constraints; and the condition [[
2
[[
2
evaluated for each point of the spectrum. To each
point in the spectrum it corresponds a [[
2
evolution, with the respective color, in Figure (A.2).
a
eq1
a
int
a
eq2
100
10
12
10
22
10
32
10
42
10
52
10
20
100
10
24
10
46
10
68
10
90
a

2
Figure A.2: Evolution of [[
2
for each mode calculated in the spectrum, from an initial moment when
the mode exits the horizon, until it reenters again.
40
1
10
10
10
20
10
30
10
40
10
50
10
60
10
16
10
13
10
10
10
7
10
4
0.1
a
d
d
a

2
r
e
l
e
r
r
o
r
Figure A.3: Relative error in the derivative of [(a)[
2
, with respect to the scale factor, for each mode
calculated in the spectrum of Fig. (A.1).
1
10
10
10
20
10
30
10
40
10
50
10
60
10
17
10
14
10
11
10
8
10
5
a

d
X
d
a
r
e
l
e
r
r
o
r
Figure A.4: Relative error in the derivative of X(a), with respect to the scale factor, for each mode
calculated in the spectrum of Fig. (A.1).
41
42
Bibliography
[1] A. Y. Kamenshchik, U. Moschella and V. Pasquier, Phys. Lett. B 511 (2001) 265 [arXiv:gr-
qc/0103004].
[2] N. Bilic, G. B. Tupper and R. D. Viollier, Phys. Lett. B 535 (2002) 17 [arXiv:astro-ph/0111325].
[3] M. C. Bento, O. Bertolami and A. A. Sen, Phys. Rev. D 66 (2002) 043507 [arXiv:gr-qc/0202064].
[4] M. Bouhmadi-Lopez, Pedro Frazao, Alfredo B. Henriques, submitted for publication [arXiv:astro-
ph/0910.5134].
[5] L. Parker, Phys. Rev. 183 (1969) 1057.
[6] B. Allen, Phys. Rev. D 37 (1988) 2078.
[7] E. Komatsu et al. [WMAP Collaboration], Astrophys. J. Suppl. 180 (2009) 330 [arXiv:0803.0547
[astro-ph]].
[8] A. H. Guth, Phys. Rev. D 23, 347 (1981).
[9] A. A. Starobinsky, Phys. Lett. B 91, 99 (1980).
[10] Edmund J. Copeland, M. Sami and Shinji Tsujikawa, Int. J. Mod. Phys. D 15 (2006) 1753 [hep-
th/0603057].
[11] R. Bean and Dore, Phys. Rev. D 68 (2003) 023515 [arXiv:astro-ph/0301308].
[12] H. Sandvik, M. Tegmark, M. Zaldarriaga and I. Waga, Phys. Rev. D 69 (2004) 123524 [arXiv:astro-
ph/0212114].
[13] L. Amendola, F. Finelli, C. Burigana and D. Carturan, JCAP 0307 (2003) 005 [arXiv:astro-
ph/0412638].
[14] M. C. Bento, O. Bertolami, N. M. C. Santos and A. A. Sen, Phys. Rev. D 71 (2005) 063501
[arXiv:astro-ph/0412638].
[15] T. Barreiro, O. Bertolami and P. Torres, Phys. Rev. D 78 (2008) 043530 [arXiv:0805.0731 [astro-
ph]].
[16] M. Bouhmadi-Lopez and J. A. Jimenez Madrid, JCAP 0505 (2005) 005 [arXiv:astro-ph/0404540].
[17] A. Kamenshchik, C. Kiefer and B. Sandhofer, Phys. Rev. D 76 (2007) 064032 [arXiv:0705.1688
[gr-qc]].
[18] M. Bouhmadi-Lopez, P. F. Gonzalez-Diaz and P. Martin-Moruno, Phys. Lett. B 659 (2008) 1-5
[arXiv:gr-qc/0612135].
[19] M. Bouhmadi-Lopez, P. F. Gonzalez-Diaz and P. Martin-Moruno, Int. J. Mod. Phys. D 17 (2008)
2269 [arXiv:0707.2390 [gr-qc]].
[20] Z. Keresztes, L. A. Gergely, V. Gorini, U. Moschella and A. Y. Kamenshchik, Phys. Rev. D 79
(2009) 083504 [arXiv:0901.2292 [gr-qc]].
[21] M. Bouhmadi-Lopez, C. Kiefer, B. Sandhofer and P. V. Moniz, Phys. Rev. D 79 (2009) 124035
[arXiv:0905.2421 [gr-qc]].
43
[22] A. R. Liddle and D. H. Lyth, Cosmological Ination and Large-Scale Structure, (Cambridge Uni-
versity Press, 2000).
[23] E. D. Stewart and D. H. Lyth, Phys. Lett. B302, 171 (1993), gr-qc/9302019.
[24] J. E. Lidsey et al., Rev. Mod. Phys. 69, 373 (1997), astro-ph/9508078.
[25] A. Kosowsky and M. S. Turner, Phys. Rev. D 52, (1995) 1739, astro-ph/9504071.
[26] L. E. Mendes, A. B. Henriques, and R. G. Moorhouse, Phys. Rev. D 52, (1995) 2083.
[27] T. L. Smith, M. Kamionkowski, and A. Cooray, Phys. Rev. D 73 (2006) 023504 [arXiv:astro-
ph/0506422].
[28] A. Buonanno, G. Sigl, G. G. Raelt, H. T. Janka and E. Muller, (2004), astro-ph/0412277.
[29] N. Seto, S. Kawamura, and T. Nakamura, Phys. Rev. Lett. 87, (2001) 221103.
[30] B. Abbott et al. (LIGO Scientic Collaboration), Rev. D 69, (2004) 122004, gr-qc/0312088.
[31] S. Larson, W. Hiscock, R. Hellings, Phys. Rev. D 62, (2000) 062001.
[32] H. B. Benaoum, arXiv:hep-th/0205140.
[33] P. F. Gonzalez-Diaz, Phys. Lett. B 562 (2003) 1 [arXiv:astro-ph/0212414].
[34] U. Debnath, A. Banerjee and S. Chakraborty, Class. Quant. Grav. 21 (2004) 5609 [arXiv:gr-
qc/0411015].
[35] L. P. Chimento and R. Lazkoz, Class. Quant. Grav. 23 (2006) 3195 [arXiv:astro-ph/0505254].
[36] I. S. Gradshteyn and I. Rhyzik, Tables of Integrals, Series and Products (Academic Press, 1994).
[37] M. Abramowitz and I Stegun, Handbook of Mathematical Functions (Dover, 1980).
[38] W. H. Kinney, arXiv:0902.1529 [astro-ph.CO].
[39] W. H. Kinney, E. W. Kolb, A. Melchiorri and A. Riotto, Phys. Rev. D 74, 023502 (2006)
[arXiv:astro-ph/0605338].
[40] W. H. Kinney, E. W. Kolb, A. Melchiorri and A. Riotto, Phys. Rev. D 78, 087302 (2008)
[arXiv:0805.2966 [astro-ph]].
[41] U. Seljak et al. (2004), astro-ph/0407372.
[42] P. Sa, A. B. Henriques, Phys. Rev. D 77 (2008) 064002 [arXiv:astro-ph/0712.2697v1].
[43] R. G. Moorhouse, A. B. Henriques, and L. E. Mendes, Phys. Rev. D 50 (1994), 2600.
44

Você também pode gostar