Nothing Special   »   [go: up one dir, main page]

Paper COSMIC MICROWAVE BACKGROUND RADIATION

Download as doc, pdf, or txt
Download as doc, pdf, or txt
You are on page 1of 13

COSMIC MICROWAVE BACKGROUND RADIATION: INSIGHTS INTO THE

EARLY UNIVERSE

ABSTRACT

The cosmic microwave background (CMB) is the name given to the radiation that continued
to exist after the Big Bang and beyond. These radiation changes have been the subject of
recent research, which has yielded significant new knowledge about the properties of our
universe. The first section of this textbook provides a succinct summary of contemporary
cosmology and its most significant accomplishments. Following that, the author moves on to
provide a comprehensive explanation of cosmological perturbation theory. Finally, the author
examines the theory of cosmic microwave background (CMB) and its recent developments.
More research is being done to study inflation as a potential cause of early oscillations. The
Boltzmann equation is responsible for controlling the evolution of CMB anisotropies, and the
total angular momentum technique is used in the process of calculating polarization
parameters. In addition to that, cosmological parameter estimations, spectral aberrations,
and the lensing of fluctuations in the cosmic microwave background are discussed. The first
thing that is covered in this textbook is an in-depth explanation of the theory of cosmic
microwave background anisotropies and polarization. Researchers and graduate students
working in this field will find this book to be an excellent resource since it includes tasks at
the end of each chapter as well as solutions to some of the questions that are asked.

Keywords: cosmic, Boltzmann, polarization, Examining

INTRODUCTION

The Cosmic Microwave Background (CMB) radiation sheds insight on the earliest beginning
of the universe, making it one of the most significant phenomena in the fields of cosmology
and astrophysics. A source of microwave radiation that is both modest and exceedingly
uniform, the cosmic microwave background (CMB) was discovered for the first time in 1964.
It includes the whole observable universe by its presence. Approximately a few hundred
thousand years after the Big Bang, it is a cosmic fossil that has retained crucial knowledge
about the universe. This information concerns the universe.

The primordial soup of hot, dense particles was the place where the cosmic microwave
background (CMB) first began to appear. Around three hundred and eighty thousand years
after the Big Bang, a significant shift in the universe known as recombination took place. As
a result of the interaction between protons and electrons, the atomic structure was rendered
inert at this moment in time, which resulted in the photons being released into space. The
cosmic microwave background (CMB), which represents the aftermath of this epoch, is a
moment in time that has been frozen in time and can be seen by us. The cosmic microwave
background, often known as the CMB, is a relic of the thermal radiation that existed in the
early cosmos. It is characterized by a temperature that is generally constant, reaching around
2.7 Kelvin. In spite of the fact that it seems to be uniform, it is really a source of vital
information on the beginnings of cosmic structures. This information comes in the form of
tiny temperature shifts, which are also referred to as anisotropies.

The study of these anisotropies has been significantly aided by the contributions made by
cosmological missions like as COBE, WMAP, and Planck. Not only does a comprehensive
investigation of the cosmic microwave background (CMB) increase our understanding of the
fundamental properties of the universe, but it also lends significant support to the inflationary
model, which is a concept that elucidates the apparent large-scale structure and uniformity of
the universe. It is possible for scientists to utilise the information that is stored in the cosmic
microwave background (CMB) to piece together the mystery of the early universe. This
information includes the age of the universe, its composition, and any general geometry that
may have existed at that time period. Missions both in space and on Earth will continue to
investigate the cosmic microwave background (CMB), which is expected to provide more
insights into the early universe and assist us in gaining a better understanding of how the
universe developed from the Big Bang to where it is today.

Radiation Spectrum of the Cosmic Microwave Background

In order to give information on the beginnings, development, and basic aspects of the
universe, the cosmic microwave background radiation spectra provide a vital insight into the
early phases of the cosmos. This view is essential for providing fundamental knowledge.
Arno Penzias and Robert Wilson made the discovery of this peculiar radiation in the year
1965. Due to the fact that it shows the weak light of the universe right after the Big Bang, it is
a one-of-a-kind instrument that may be used to investigate the early universe.

Revealing the Sources

In order to get a comprehensive comprehension of the CMB spectrum, it is necessary to do


study into the conditions that gave rise to this radiation. The cosmos started to cool down and
expand after the Big Bang, which led to the production of neutral hydrogen atoms via the
merging of protons and electrons. This occurred as a consequence of the universe expanding
and cooling. It was at this period of time, which is known as the recombination epoch, that
photons were finally given the opportunity to move across space without being constantly
scattered. As a result of the expansion of the universe, these primordial photons have been
twisted to microwave wavelengths, and as a consequence, they are the source of the cosmic
microwave background radiation (CMB) that humans are able to see in the current day.

The Big Bang's Thermal Relic

As an example of a thermal artefact, the spectrum of the cosmic microwave background


(CMB) is created during the early period of the cosmos, which was characterised by very
high temperatures and high levels of energy. As a result of the fact that it is a blackbody, the
radiation that it generates has a spectrum that is comparable to that of a body that is flawless
and idealised. Through the disclosure of crucial data on the temperature of the cosmos at the
moment of recombination, scientists are able to restrict cosmological possibilities. This is
made possible by the structure and intensity of the spectrum, which is distinctive to the
spectrum.

Anisotropies in the Universe's Harmony

If one observes the cosmic microwave background (CMB) from a considerable distance, it
seems to be quite flat. On the other hand, studies of temperature changes at the microscopic
level, which are referred to as anisotropies, show that the universe does in fact have some
structure. These variations, which were left behind by quantum fluctuations in the early
cosmos, provide essential information about the geometry, development, and composition of
the universe thanks to the information they provide. The quantum fluctuations were
responsible for leaving behind these variances. Many cutting-edge sensors, including as the
Planck satellite, have methodically followed these anisotropies across the sky in order to
increase our grasp of the cosmos. This has allowed us to acquire a deeper comprehension of
the universe.

Polarization Goes Beyond Temperature

The CMB spectrum not only offers information about changes in temperature, but it also
provides information about polarization. With the aid of polarization patterns in the cosmic
microwave background (CMB), which provide further insights into the characteristics of the
universe, scientists are able to research issues such as cosmic inflation, the presence of dark
matter, and the timing of reionization. These topics are included in the list of topics that may
be investigated.

As a consequence of advancements in observational technology and the implementation of


more complex experiments, our understanding of the spectrum of the cosmic microwave
background is continuously expanding. There is a possibility that the endeavour to
understand the cosmic microwave background spectrum may provide ongoing discoveries
into the early years of the universe and the rich cosmic symphony that it contains. Both
satellite missions and ground-based observatories are working towards the successful
completion of this objective.

OBJECTIVE OF THE STUDY

1. To examine the Understanding the origins of the universe via the cosmic microwave
background radiation.

2. To examine the Future and Present Research Directions in CMB Studies.

RESEARCH METHODOLOGY

Lasenby and Hobson summarise the findings and characteristics of a number of


investigations that were conducted during the years 1994 and 1996 in their paper. These
studies were conducted in the United States. The major emphasis of our attention is on the
recent findings that have been obtained from a variety of research that have made substantial
advancements since then, namely within the span of the last year. The critical parameters that
have been discovered in a number of recent studies are presented in a straightforward manner
in Table 1, which gives a handy technique for doing so.

The satellite COBE

In order to carry out measurements of primordial anisotropy, the DMR experiment makes use
of a total of six differential microwave radiometers, two of which are running at each
frequency of 31.5 GHz, 53.0 GHz, and 90.0 GHz. In the first set of COBE observations, solid
statistical evidence was obtained that demonstrated the existence of CMB changes.
Unfortunately, it was not possible to see distinguishing properties of the Cosmic Microwave
Background (CMB) at the same scale as the beam size in the DMR maps. This limited the
scope of the investigation. This was owing to the fact that even after all of the photos were
merged, the level of noise per beam zone stayed below one since it was around 45 degrees
Kelvin, and the signal to noise ratio remained below one.

The results of the investigation of the whole of the DMR data set, which covered a period of
four years, are now available to the public. In Bennett et al. (1996), you can find a detailed
collection of all of the discoveries that were discovered. There is a possibility that the data
may be used to restrict the normalization of a power law primordial spectrum on a statistical
basis. Normalization is often shown as the proposed amplitude of the quadrupole component
of the power spectrum, denoted by the symbol C2, according to a certain slope, denoted by n.

Table 1. A few recent observations of CMB anisotropy

Beam throw / beam


Trial Sort ν (GHz)
width

DMR-COBE Satellites 7.2 31.1, 53.0, 90.3

Tenerife Grounds 5.4/8.2 10.7, 15.3, 33.2

MIT/FIRS Balloons 3.7 180.0 + 3.1 high

ACME/HEMT Grounds 1.6/2.2 30.1/40.3

MAX Balloons 0.6/1.1 110.1, 180.0, 270.2

MSAM Balloons 0.4/0.7 180.5 + 3.8 high

White dish Grounds 0.19/0.46 90.1

Python Grounds 0.76/2.74 90.0

Saskatoon Grounds 1.4/2.46 26.1 – 36.2

ARGO Balloons 0.7/1.2 150.0 + 3.0 high

CAT Grounds 0.24 13.1 – 16.0

OVRO Grounds 0.2-0.5 14.4 + 32.0

IAC-Bartol
Grounds 2.0 91.1 – 272.3
OVRO

The individual anisotropy patterns that are included inside the beam size-scale maps are
beginning to have statistical significance now that four years' worth of data has been
amassed. Figure 11 displays the all-sky maps that Bennett et al. created at each frequency. On
account of the fact that the signal-to-noise ratio in these regions is around 2 sigma per 10-
degree sky patch at the moment, it is predicted that some of the features in these maps that
are situated at a considerable distance from the Galactic plane will be actual fluctuations in
the cosmic microwave background. Characteristics that are constant across all of the
frequencies may now be seen with relative ease.

The trials conducted in Tenerife


These experiments were described in great detail in Lasenby and Hancock. In the next part,
we will briefly go over the details that are most significant. The Tenerife experiments consist
of three different devices that operate at 10, 15, and 33 GHz. These devices were created and
manufactured at Jodrell Bank, and they are operating on the island of Tenerife, which is
managed by the IAC. Drift scanning in the right ascension at a fixed declination and sampling
at intervals of a beamwidth in the declination are both methods that are used in order to
collect information via the use of drift scanning.

There is the possibility of constructing a two-dimensional sky map that is entirely sampled at
each frequency. The low frequency surveys that were carried out by Haslam et al. and Reich
and Reich indicate that there is a minimum in Galactic foreground emission. As a result, the
first observations were concentrated on the sky strip at a declination of +40.0 degrees.

Tenerife is now engaged in an endeavor with the objective of mapping about 4,000 square
degrees of sky at frequencies of 10, 15, and 33 GHz. These research, in conjunction with the
COBE data collection that will take place over the course of four years, will continue to
provide a great resource for large-scale CMB anisotropy measurements, which will enable
cosmological hypotheses to be directly evaluated. A fresh estimate of the fluctuation
amplitude for values that are quite near to -20 has been generated by using the data from
December 40◦. This estimation takes into account an atmospheric component that was not
taken into account in the first study conducted by Hancock et al. A comparison of this with
theoretical curves is shown below.

The Radio Observatory in Owen's Valley

The Owen's Valley Radio Observatory (OVRO) is a single-dish antenna telescope with a
diameter of forty metres that is situated on the ground. The sensitivity of the telescope has
been improved as a result of the recent installation of a HEMT receiver that is capable of
operating at two different frequencies. Within the scope of the first experiment, there were
significant constraints placed on the many possibilities for the creation of galaxies.
Figure 1. Saskatoon 3-year graphic comparing the area under analysis to the whole sky
coverage of COBE

DATA ANALYSIS

The exponential development in the amount of data that can be accessed from CMB research
is causing the necessity of improving analytical tools to increase in parallel with this growth.
It ought to be possible to exclude foreground information and leave behind a "clean" CMB
map by using high-precision multifrequency observations of the same sky patch (the satellite
will cover the whole sky). According to the findings of a number of experiments, Maximum
Entropy is one strategy that has shown to be very effective throughout the course of time.

Assuming that there is a hypothesis H and some data D, Bayes' theorem states that the
posterior probability is equal to the sum of the prior probability and the evidence Pr(D), with
the prior probability being normalized by the evidence.

If the probability distribution of the instrumental noise follows a Gaussian distribution on


each frequency channel, then the distribution is said to be a multivariate Gaussian. When we
make the assumption that the anticipated value of the noise is zero at each and every
frequency that is detected, we are able to compute the probability by using the following
formula:

Methods that are considered to be conventional for maximum entropy presume that the image
H has a positive additive distribution (PAD). Nevertheless, the MEM approach may be
expanded to include photos with positive and negative values by seeing them as the
difference between two PADS. This makes it possible for photographs to be analyzed.

where the letter U stands for the positive side of H and the letter V stands for the negative
aspect. Taking into consideration this circumstance, the cross entropy may be represented
when

Six different input maps were used by the researchers before the addition of Gaussian noise to
each frequency. These maps included the cosmic microwave background (CMB), thermal and
kinetic SZ, dust emission, free-free emission, and synchrotron emission. Following the
application of MEM with the Bayesian value for α and the implementation of the procedure
using the average power spectra of each channel, the features in all six maps were
successfully recovered.
The kinetic SZ was found to be unrecoverable, despite the fact that all of the other power
spectrum data had been retrieved to some degree. This was discovered in the absence of any
preceding power spectrum data. To be more specific, the CMB and dust exhibited residual
errors of 6µK and 2µK per pixel, respectively, which rendered them almost indistinguishable
from the input maps. To illustrate the difference between the output from MEM and the input
maps for the situation in which the average power spectrum is taken into consideration,
Figure 2 is shown. Not only can MEM properly replicate the Gaussian CMB, but it also
successfully recreates the non-Gaussian thermal SZ effect. This is something that is
immediately obvious.

Figure 2. To the left of the picture are the input maps that were used in the Planck
simulations to model the thermal SZ effect and the CMB using a CDM. As you can see
on the right, the MEM reconstructions are shown as well.

The power spectrum of CMB in relation to experimental points

The preceding sections ought to have made it abundantly evident that the CMB data are
coming closer and closer to the point where the shape and normalisation of the power
spectrum may be compared with theory and predictions. The scale combination that they give
is great for beginning to trace the structure of the first Doppler peak, particularly with the
recently published data from CAT and Saskatoon.

In addition to being improperly calibrated and noisy, the current data from the Cosmic
Microwave Background (CMB) will also include residual pollution from the Galaxy or
individual radio emitters, or maybe both. This is something that should be taken into
consideration. Despite the best efforts of experimenters to avoid or reduce these effects, the
process of producing really "clean" CMB data, which is free of these effects to a certain
degree of certainty, is still in its early stages.

Second, the accuracy of the estimated parameters in any theory-data comparison is


determined by the quality of the theoretical models and assumptions that are used as the basis
for the comparison. In the event that it is decided that the theory that replaces CDM is not a
practical choice, for example, the constraints on Ω that were established below will need to
be reassessed. Due to the fact that many of the components that comprise the power spectrum
are not theory-specific, it is anticipated that some of the results will not be much impacted.
This includes the physics of recombination, which is only dependent on atomic physics when
it has been well comprehended.

Figure 3. Power spectrum analytic fit vs experimental points


Figure 4. This image depicts the universe's history from the big bang forward.

Previous Actions

In accordance with what was said before, it is presumed that the traditional, spatially flat,
ΛCDM model is the background cosmological model. The model may be defined by a total
of four parameters. It is the latter two variables that are responsible for representing the sound
horizon, while the first two variables are responsible for indicating the actual baryon and
CDM densities (with h being tied to the Hubble parameter). Table 2 is located farther down
on this page, and it contains the priors for these four parameters.

Table 2. The precedence values of the four parameters characterizing the spatially flat
background ΓCDM model

Parameter background Minimum threshold Maximum amount

Ωb h 2 0.006 0.3

Ωc h 2 0.02 0.98

Θ 0.6 10.1

τ 0.03 0.7
When selecting the priors for the two inflationary models, the selection process is carried out
in such a manner as to reduce the duration of inflation, generate spectral indices that are in
accordance with the criteria, and maintain the amplitude of the resulting scalar spectra in
close proximity to the COBE value. There is a display of the priors for inflationary models in
Table 3.

Table 3. In the power law scenario, the two inflationary potentials and priors on the
three primordial spectra parameters are relevant.

Example Minimum threshold Maximum amount

2.6 4.1

Power-law situation 0.4 1.6

0.1 1.1

-0.76 -0.57

0.0 2 x 10.0
Model of chaos with
modulation in sinusoids
2.0 x 10.0 1.0

− *

0.6 1.26

1.2 × 10.1 2 x 1.8


Model of Axion
Monodromic
3.0 × 10.2 1 x 10.3

− *

The spectra and enhanced fit quality

The parameter χ 2 eff, which represents the least squares, is provided in Table 4 for each of
the models and datasets that were considered in this study. It is evident from the table that the
monodromic model provides a better fit to the data by around 5-6 standard deviations, which
is referred to as the χ 2 eff. There seem to be two other spots that are suggested by the table as
well. Despite the fact that the chaotic model with sinusoidal modulation does not perform as
well as the monodromic model, the addition of small-scale data from ACT enhances the
performance of the model, which indicates that the data may favour oscillations.

Second, the data tends to support potential oscillations with constant amplitudes, such as in
the monodromy model, over chaotic oscillations with variable amplitudes, such as in the
chaotic model with sinusoidal modulations. This is especially true when compared to the
chaotic model with sinusoidal modulations. In point of fact, this gives credibility to previous
discoveries that verify this, namely that the data are far better characterised by trans-
Planckian scale oscillations in the primordial spectrum of a certain amplitude. At this point, it
is quite interesting to investigate if there are any localised multipole windows in which the fit
is improved.

Table 4. At the low multipoles, the χ 2 eff for the CMB TT spectrum has been computed
using the Gibbs method in the WMAP likelihood code.

Collections of Data
MAP-7.1 MAP-7.1 + ACT3.1
Example

Power-law situation 7468.4 7500.5

Model of chaos with


7467.5 7498.1
modulation in sinusoids

Model of Axion
7462.0 7495.3
Monodromic

All of the models, including the power law example, have their TE amplitudes and the
accompanying CMB EE angular power spectra shown in Figure 5.

Figure 5. The difference in χ 2 eff with respect to the reference model, i.e. ∆χ 2 eff = [χ 2
eff (model) − χ 2 eff (power law)

CONCLUSION

Several speculative possibilities strongly encourage the hunt for faint statistical isotropy
breakdowns in the cosmos. The cosmic microwave background variations are one of the most
encouraging observational probes of the universe's SI. Possible sources of the SI violation
include the anisotropic primordial power spectrum and the fluctuations of the baryon-photon
fluid at final scattering. Earlier it was said that some of the abnormalities in CMB maps may
be explained by changes to the usual physics at the surface of last scattering. In an
inflationary model, the step's introduction may be seen as a sudden shift in a possible
parameter. However, it is true that it is quite haphazard, and more investigation into feature
creation and the enhancement of fit in more appropriately driven inflationary models is
required. This is made possible by two field models. For example, the two field models may
readily provide a temporary break from sluggish roll inflation with appropriately selected
parameters. An inflationary era, when added to the traditional hot big bang scenario, explains
the physical events that occurred in the early cosmos. A study of the universe's overall
characteristics has been made possible by basic cosmological measurements made possible
by high precision cosmology observations in the past and in the present. When compared to
other current, reasonable theories about the early cosmos, these findings substantially support
the conventional cosmological model.

REFERENCES

1. H. K. Eriksen, F. K. Hansen, A. J. Banday, K. M. Gorski and P. B. Lilje, Astrophysical


Journal 605, 14 (2004).

2. J. Dunkley et al., Astrophys. J. Suppl. 180, 306 (2009); E. Komatsu et al., Astrophys.
J. Suppl. 180, 330 (2009).

3. J. Bardeen, Physical Review D 22, 1882 (1980); J. M. Stewart, Class. Quantum Grav.
7, 1169 (1990).

4. L. P. Grishchuk, Sov. Phys. JETP 40, 409 (1974); A. A. Starobinsky, Sov. Phys. JETP
Lett. 30, 682 (1979).

5. A. Papoulis and S. Unnikrishna Pillai, Probability, Random Variables and Stochastic


Processes.

6. B. D. Wandelt, D.L. Larson and A. Lakshminarayanan, Physical Review D, 70,


083511, (2004).

7. M. Bridges, F. Feroz, M. P. Hobson and A. N. Lasenby, Mon. Not. Roy. Astron. Soc.
400, 1075 (2009); H. V. Peiris and L. Verde, Physical Review D 81, 021302 (2010);
Z.-K. Guo, D. J. Schwarz and Y. -Z. Zhang, JCAP, 08 031 (2011).

8. M. J. Mortonson, C. Dvorkin, H. V. Peiris and W. Hu, Physical Review D 79, 103519


(2009).

9. R. K. Jain, P. Chingangbam, L. Sriramkumar and T. Souradeep, Physical Review D


82, 023509 (2010).

10. W. H. Kinney and K. T. Mahanthappa, Physical Review D 52, 5529 (1995); Physics
Letters B 383, 24 (1996).

11. W. H. Press, S. A. Teukolsky, W. T. Vetterling and B. P. Flannery, Numerical Recipes


in FORTRAN 90, Second edition (Cambridge University Press, Cambridge, England,
1996).

12. C. Gordon, D. Wands, B. A. Bassett and R. Martens’, Physical Review D 63, 023506
(2000).
13. P. Hunt and S. Sarkar, Physical Review D 70, 103518 (2004); Physical Review D 76,
123504 (2007).

14. X. Chen, R. Easther and E. A. Lim, JCAP 0706, 023 (2007); JCAP 0804, 010 (2008).

15. M. Aich, D. K. Hazra, L. Sriramkumar and T. Souradeep, arXiv:1106.2798 (2011).

16. C. Pahud, A. R. Liddle, P. Mukherjee and D. Parkinson, Physical Review D 73,


123524 (2006).

17. E. Komatsu, J. Dunkley, M. R. Nolta, C. L. Bennett, B. Gold, G. Hinshaw, N. Jarosik,


D. Larson, M. Limon, L. Page, et al., Astrophysical Journal, 180, 330 (2009).

18. A. de Oliveira-Costa, M. Tegmark, M. Zaldarriaga and A. Hamilton, Physical Review


D 69,063516 (2004).

You might also like