12.01.2015 Views

Cosmology with CMB Anisotropy - iucaa

Cosmology with CMB Anisotropy - iucaa

Cosmology with CMB Anisotropy - iucaa

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

¡£¢¥¤§¦©¨¢ £¡¦<br />

¨ £¢"!#§$%&§ '(¢)¡ §*+¢,- .<br />

¨$0/(132¨£¡ 456¨2876¦9¦©%<br />

: ; = @BAC


Certificate of the Guide<br />

CERTIFIED that the work incorporated in the thesis "<strong>Cosmology</strong> <strong>with</strong> <strong>CMB</strong> <strong>Anisotropy</strong>"<br />

Submitted by Mr. Amir Hajian Forushani was carried out by the candidate under my<br />

supervision/guidance. Such material as has been obtained from other sources has been<br />

duly acknowledged in the thesis.<br />

Prof. Tarun Souradeep<br />

(Thesis Supervisor)<br />

Date:<br />

iii


Declaration<br />

I declare that the thesis entitled "<strong>Cosmology</strong> <strong>with</strong> <strong>CMB</strong> <strong>Anisotropy</strong>" submitted by<br />

me for the degree of Doctor of Philosophy is the record of work carried out by me<br />

during the period from January 2002 to January 2006 under the guidance of Prof.<br />

Tarun Souradeep and has not formed the basis for the award of any degree, diploma,<br />

associateship, fellowship, titles in this or any other University or other institution of<br />

Higher learning.<br />

I further declare that the material obtained from other sources has been duly acknowledged<br />

in the thesis.<br />

Amir Hajian Forushani<br />

Date:<br />

v


I gratefully dedicate this thesis to my parents,<br />

whose love and support I will forever be thankful.<br />

And to my wife, Sanaz,<br />

for her love, support, encouragement and for giving me a happy and<br />

complete life.<br />

vii


Abstract<br />

One of the main challenges in the study of the Cosmic Microwave Background (<strong>CMB</strong>)<br />

anisotropy is to develop methods to extract maximum information from the rich source<br />

of information provided by this random field on the sky. When the condition of statistical<br />

isotropy (SI) holds, the widely used angular power spectrum has proved to be a very<br />

useful quantity which contains the whole information encoded in the <strong>CMB</strong> anisotropy<br />

field. However, there are certain mechanisms that violate the SI condition. This thesis<br />

is dedicated to the study of this fundamental assumption in <strong>CMB</strong> studies. We have<br />

formulated the problem mathematically, constructed a quantitative method to describe<br />

SI violation (bipolar power spectrum), studied number of sources of violation of SI<br />

condition and finally examined this fundamental assumption using the most recent full<br />

sky <strong>CMB</strong> anisotropy maps provided by the first-year data of Wilkinson Microwave<br />

<strong>Anisotropy</strong> Probe (WMAP). The main conclusions are:<br />

• Statistical isotropy of <strong>CMB</strong> anisotropy is an important property of <strong>CMB</strong> anisotropy<br />

field which can be quantified and checked in distinct from Gaussianity of the field.<br />

Bipolar Power Spectrum (BiPS) is an orientation independent quantity which<br />

is sensitive to structures and patterns in the underlying two-point correlation function<br />

and is a very useful tool for searching for departures from SI.<br />

• We find no strong evidence for SI violation in the full sky <strong>CMB</strong> anisotropy maps<br />

based on the first-year data of WMAP up to l ∼ 60 − 70.<br />

• The possibility of compact spaces in cosmic topology is a theoretically well motivated<br />

possibility that has also been observationally targeted. The cleanest signaix


ture of the topology of the universe is written on the microwave sky. The existence<br />

of preferred directions along the highly correlated spots is the smoking gun of compact<br />

spaces and can be exploited to detect them. We propose BiPS as a fast and<br />

orientation independent method to look for such signatures in <strong>CMB</strong> anisotropy<br />

map.<br />

• Hiding an anisotropic template such as those of Bianchi models in a ΛCDM <strong>CMB</strong><br />

anisotropy temperature map, will lead to observable effects on bipolar power spectrum.<br />

The null BiPS of WMAP can be used to put tight constraints on the<br />

existence of such hidden patterns in the WMAP data.<br />

• Observational artifacts such as non-circular beam, inhomogeneous noise correlation,<br />

residual striping patterns and foreground residuals are potential sources of SI<br />

breakdown. Our null BiPS results confirm that these artifacts do not significantly<br />

contribute to the first-year data of WMAP up to l ∼ 60 − 70.<br />

x


List of Publications<br />

This thesis is based on the following publications.<br />

Chapter 2: A. Hajian, T. Souradeep, “Measuring Statistical Isotropy of <strong>CMB</strong>”,<br />

Astrophys.J. 597 (2003) L5-L8.<br />

A. Hajian, T. Souradeep, “The Cosmic Microwave Background Bipolar Power<br />

Spectrum: Basic Formalism and Applications” (astro-ph/0501001) (Accepted for<br />

publication in ApJ).<br />

T. Souradeep, A. Hajian, “Statistical Isotropy of Cosmic Microwave Background<br />

Radiation <strong>Anisotropy</strong>”, Pramana, vol. 62, no. 3, 793-796; 2004.<br />

Chapter 3: A. Hajian, T. Souradeep, “Statistical Isotropy of <strong>CMB</strong> and Cosmic<br />

Topology”(astro-ph/0301590), to be submitted to PRL.<br />

Chapter 4: A. Hajian, D. Pogosyan, T. Souradeep, C. Contaldi and R. Bond, 2004<br />

in preparation; Proc. 20th IAP Colloquium on Cosmic Microwave Background<br />

physics and observation, 2004.<br />

Chapter 5: T. Ghosh, T. Souradeep and A. Hajian, 2006, in preparation.<br />

Chapter 6: R. Saha, A. Hajian, T. Souradeep and P. Jain, 2006, in preparation.<br />

Chapter 7: A. Hajian, T. Souradeep, N. Cornish, “Statistical Isotropy in the WMAP<br />

Data: A Bipolar Power Spectrum Analysis”, Astrophys.J. 618 (2004) L63-L66.<br />

T. Souradeep, A. Hajian, “Statistical isotropy of <strong>CMB</strong> anisotropy from WMAP”,<br />

xi


Proc. of General relativity and Gravitation (JGRG-14), Yukawa Institute for<br />

Theoretical Physics, Nov 29-Dec 3, 2004.<br />

During my PhD at IUCAA I also worked on another project, (J.V. Narlikar, R.G. Vishwakarma,<br />

A. Hajian, T. Souradeep, G. Burbidge, F. Hoyle, “Inhomogeneities in the<br />

Microwave Background Radiation interpreted <strong>with</strong>in the framework of the Quasi-Steady<br />

State <strong>Cosmology</strong>”, Astrophys.J. 585 (2003) 1-11), but that research is not included in<br />

this thesis.<br />

xii


Acknowledgements<br />

In constant action are the clouds, wind, sun, and universe,<br />

So acknowledge your sustenance upon savoring each morsel.<br />

Golestan (The Rose Garden)<br />

Sheikh Mosleheddin Sádi Shirazi (1210-1290)<br />

This thesis, although is formally a report of my work during the last four years, is in<br />

fact a result of collective efforts and help from many people in the past 27 years. In<br />

the beginning I would like to thank everyone who have helped me through out my life.<br />

I wish to thank those whose support, help and encouragement have been invaluable to<br />

me, although I may not name all.<br />

First and foremost I should start by thanking my parents, to whom I dedicate this<br />

thesis. During the long difficult years of war and in the poverty induced by revolution,<br />

imposed war, international sanctions and embargo against Iran, my parents worked hard<br />

and sacrificed their future, to protect us and give us the best one could have in those<br />

years. They taught us to work hard, aim high and not to give up even if everything<br />

seems to be going wrong. Their support constantly continued till now. They put their<br />

house in pledge to get me exempted from the military service when I wanted to go out of<br />

Iran to continue my studies. I am forever grateful to them for every bit of their support<br />

and care all through my life.<br />

I owe special thanks to my brother and sister, Hamid and Zahra, who have always<br />

been there for me at all the important times of my life. Hamid has always inspired me<br />

xiii


through his passion for science, and showed me there are much more things to learn<br />

than I thought. He has been a wonderful brother and a great supporter for me. Zahra<br />

has given me every reason to live happily. Her vitality and energy has brought a lot of<br />

joy to our family. I hope and pray for the happiness and success of both of them. I<br />

am also grateful to my grandmother for seeing me and helping me in my life specially<br />

during my study at Sharif. She gave me the best books I have ever read. “Principles of<br />

Thermodynamics” that she gifted me changed my life and helped me realize my passion<br />

for physics in the early years of high school.<br />

I believe the credit of the beginning of my interest in research belongs to my wonderful<br />

high school, the National Organization for Development of Exceptional Talents. The<br />

infinite degree of freedom they gave us, the interesting workshops, the exciting research<br />

projects and very good teachers are the major things I can remember from my high<br />

school days. I am specially indebted to Mr. Ejei and two of my teachers Mr. Razi Zadeh<br />

and Mr. Mollabashi for providing us an excellent place to live, learn and experiment.<br />

I owe great thanks to Houshang Karimi whose excellent courses on electromagnetism<br />

and calculus introduced me to a new phase of learning physics and Ali Sedighi <strong>with</strong><br />

whom I shared my fascinations and joys of early days of adventuring <strong>with</strong> physics. In<br />

my final year at high school, the first students’ conference of physics was organized in<br />

Sanandaj. For many years, this conference gave me great motivation to study physics.<br />

I was specially impressed by Ali Nayeri and his lecture on the fractal universe. He has<br />

always inspired me and answered my mails very kindly. I remember that he discovered<br />

IUCAA and told us about that when we were very young.<br />

From my undergraduate years I am most of all indebted to Yousef Sobouti, director<br />

of the Institute for Advanced Studies in Basic Sciences. His idea of giving everyone<br />

a chance to do well has helped many students including me. The unique scientific<br />

atmosphere of IASBS gave me a great opportunity to learn from everyone. Specially<br />

I learned a great deal from my Masters’ thesis supervisor Amir H. Fathollahi, a rolemodel<br />

both as a man and a very bright scientist. Two years at IASBS were perhaps the<br />

happiest of my life. I am specially grateful to my undergraduate friends Mohammad<br />

Charsooghi, Negar Aliakbarzadeh, Farshid Mohammad Rafiee, Abouzar Shirzad, Ali<br />

Bagheri Nejad, Ali Tabeai, Afshin Khezerloo, Said Paktinat, Mohammad Ashoori and<br />

other classmates at Sharif and IASBS who remained very good friends of mine for a<br />

xiv


long time and over long distances.<br />

When I was at IASBS, Ajit Kembhavi visited us and invited me to visit IUCAA<br />

for the first time. That was my introduction to IUCAA. He has always impressed me<br />

<strong>with</strong> the care, attention and hospitality. Considering my time in India, which is most<br />

relevant to this thesis, first and foremost I would like to thank my thesis supervisor<br />

Tarun Souradeep for his constant support, encouragement and caring during the whole<br />

period of my PhD. He was (and remains) my only reason to move to India and the<br />

greatest credit of every success that I might have had in these years belongs to him and<br />

his wonderful guidance. Tarun was much more than just a thesis supervisor for me.<br />

He knew when to let me struggle alone and when to help me out of the situations. He<br />

introduced me to all the right people and very patiently helped me step by step to gain<br />

exposure in the field. I should never forget the great help when I was absolutely stuck<br />

for my registration at the university. His heartwarming words helped me a lot during<br />

frustrations of the early years. I am specially thankful to him for the belief he showed<br />

in me even when I made terrible mistakes. When I made my first <strong>CMB</strong> map, he gave<br />

it much more credit than it probably deserved. I offer my deepest gratitude to him and<br />

his family, Sucheta and Sourath, for their warm hospitality in all these years. I shall<br />

never forget memorable moments I spent at their house, and may be I have never told<br />

them how much our dinner conversations, picnics, Tarun’s readings of Indian mythology,<br />

Sucheta’s advices, Sourath’s poems and songs and every moment of being <strong>with</strong> them<br />

meant to me. Upon my arrival at IUCAA, Jayant Narlikar welcomed me very warmly<br />

and invited me to his home. He, as the director of IUCAA, supported me very strongly<br />

during the first two years of my PhD and showed me a way out of several complicated<br />

bureaucratic situations. When my visa took time and I couldn’t make it for one term<br />

at the graduate school, he gave me the chance to study on my own and just give the<br />

exams. It was a great experience for me, at both intellectual and personal levels, to<br />

communicate <strong>with</strong> him and his wife. I should thank IUCAA for providing me a calm<br />

atmosphere to work. I am specially thankful to the director of IUCAA, Naresh Dadhich,<br />

for his kind support during the last couple of years. I will always remember the night<br />

he invited all Iranian students to his house to celebrate the occasion of winning Nobel<br />

Peace Prize by Shirin Ebadi. Many IUCAA staff helped me in these years. I wish to<br />

thank all of them. Neelima Pagonkar always helped me <strong>with</strong> the paper work, specially<br />

xv


when I needed a long leave for my wedding. Susan Kuriakose was not only helpful in the<br />

official work but also a very good friend of ours. Rajesh Pardeshi saved me from missing<br />

many deadlines by his speed in getting things done. Lata Shankar <strong>with</strong> her ability in<br />

planning a trip helped me a lot in my long pre-doctoral journey. I am also thankful<br />

to many other friends at IUCAA who helped me in these four years. Anand Sengupta<br />

helped me out of many Linux-related troubles and gave me his thesis template. Sanjit<br />

Mitra was a very nice officemate. Subharti Ray cured my Microsoft windows several<br />

times and helped me in many situations. Neeraj Gupta came to my rescue when I<br />

had serious troubles understanding the ISM. He also helped me a lot <strong>with</strong> AIPS during<br />

our Radio Astronomy course. I enjoyed working <strong>with</strong> him on cleaning maps and the<br />

memorable neutron star hunting experience during the graduate school. Ayesha Begum<br />

let me pay the SEVIS fees <strong>with</strong> her credit card when no bank gave foreigners a credit<br />

card in India. Mahdi Bazargan saved us from cold nights of IUCAA. Ali Dariush <strong>with</strong><br />

his extreme vitality brought joys to my life for the short time he stayed at IUCAA.<br />

Arman Shafieloo introduced me to interesting people. Meysam Namayandeh gave me a<br />

thorough example of a trust worthy friend, I am very happy he finally recovered from the<br />

dengue fever. Elisabeth, Maelys, Ondine and Emannuel Rollinde provided a very warm<br />

family atmosphere for me during the short time I knew them, my first joyful experiences<br />

of baby sitting dates back to the days they were <strong>with</strong> us. Anil, Nikita and Mammad<br />

Sandwichi helped us in their own ways. Vida Rahiminejad took the trouble of helping<br />

me through the paper work in my absence and has always been a very good friend of<br />

ours. My very special thanks belongs to Dr. Mohammad Kafi, science counselor and<br />

director of science, technology and education in the Embassy of I.R. of Iran in New<br />

Delhi. In very difficult situations he helped and supported me. I am forever grateful to<br />

him for his invaluable supports and encouragements which I will never forget.<br />

Being an Iranian student in India I always had the trouble of finding funds for my<br />

travel. Tarun always helped me greatly to find kind people to fund me. I am deeply<br />

grateful to all of them for their support and hospitality which allowed me meet brilliant<br />

people and learn many things from them in my trips. In my first year Norma Sanchez<br />

covered all my expenses to attend the Chalonge School. I can never forget the excitement<br />

of being there. Farhad Ardalan invited me to give a talk on WMAP at the Institute<br />

for Theoretical Physics and Mathematics (IPM). In that trip I met the young society<br />

xvi


of cosmologists of Iran and it helped me to strengthen my ties to my home country. In<br />

my predoctoral trip I met Simon Prunet very briefly, but in the short time I learned<br />

a lot from him. The Chapter on foregrounds in my thesis is mostly the result of my<br />

discussion <strong>with</strong> him. He also helped me out when I was stuck <strong>with</strong> the computation<br />

of Clebsch-Gordan coefficients. Mr and Mrs Rollinde let me stay <strong>with</strong> them during<br />

my visit to Paris and took care of me like their own son. I am extremely grateful to<br />

them and I don’t know what I could have done in Paris if they did not accommodate<br />

me in their house. At ICTP Seif Randjbar-Daemi and Uros Seljak hosted me very<br />

kindly. Maria Jesus Pons helped me a great deal to attend the school on data analysis<br />

in cosmology which was very useful for me. Andrea Tartari and Georgio Sironi hosted<br />

me in Milan where for the first time I could see the horns <strong>with</strong> which <strong>CMB</strong> measurments<br />

are made (they were much smaller than what I thought!). Their hospitality was superb<br />

and Milan is always a memorable city for me since then. At Oxford I enjoyed very warm<br />

hospitality of Joe Silk, Sadegh Kouchfar and Asim Mahmood. My visit to Oxford was<br />

very fruitful specially because the Bianchi models were hot at that time and I could<br />

learn a lot about them from Joe Silk and Pedro Ferrera. Our discussions gave me the<br />

idea of what I later did on Bianchi models. It was a great experience to work at CITA<br />

and I am greatly indebtful to Dick Bond for inviting me and giving me full support<br />

to be there for a long period. Nazli and Houshang Karimi made my stay at Toronto<br />

wonderful. I can never thank them enough for their warm and kind Iranian hospitality.<br />

I am also thankful to Subhabrata Majumdar, Faranak and Yaser Kerachian for their<br />

kind help and hospitality. My other great experience was the period I spent <strong>with</strong> Dmitry<br />

Pogosyan. It was a wonderful time of excitingly hard work. Dmitry accommodated me<br />

at his own house. That was a great opportunity for me to discuss all through the day<br />

and night and learn a lot from him at every moment. He gave me the co-added noise<br />

covariance matrix of WMAP and the work I did on Poincare Dodecahedral Spaces and<br />

also the BiPS of anisotropic noise was done during this period. My trip to Harvard was<br />

very useful for me for which I am indebted to Niayesh Afshordi. I enjoyed discussions<br />

<strong>with</strong> Niayesh, Ramesh Narayan and Avi Loeb. I am also thankful to Banibrata, Ghazal<br />

and Niayesh for their kind hospitality. At UC Davis Lloyd Knox hosted me and I had a<br />

great chance of learning from him. He helped me understand some issues on statistics<br />

of power spectrum and also taught me how to write the equations in powerpoint in<br />

xvii


color! My greatest experience was visiting Princeton. I enjoyed discussions <strong>with</strong> David<br />

Spergel, Lyman Page and Olivier Dore. For a long time I had worked on the WMAP<br />

data and it was a thrill for me to present my work for the WMAP team members. I am<br />

also thankful to Lyman Page and David Spergel who helped me in getting my US visa<br />

and moving to Princeton. At Princeton I enjoyed very warm hospitality of Mojgan and<br />

Hormoz. They made me feel at home for the whole period of my stay and I will never<br />

forget the great fun and good memories of the cold evenings of New Jersey warmed <strong>with</strong><br />

their kind Iranian hospitality.<br />

During four years of my PhD, Yahoo Messenger was my only way of remaining in<br />

touch <strong>with</strong> friends and family. I wish to thank Yahoo for such a wonderful facility and<br />

also for Yahoo Mail which is the most reliable mail server I have ever seen. Blogger<br />

gave me the possibility of writing, reading and publishing in Persian. This, along <strong>with</strong><br />

Orkut, Gmail and Google search have been great inventions of Google and I can not<br />

express how much they have affected my life. All these were impossible to use if I did<br />

not have Windows XP which makes life very easy, allows you to use voice, video and<br />

graphics and lets you read and write in your own script.<br />

Last, but by no means least, my love, admiration, and gratitude goes to my wife,<br />

Sanaz. During the last three years, she has always been my closest and best friend. She<br />

came into my life in a terrible time and protected me in all the dark periods by her<br />

precious love. Sanaz bore <strong>with</strong> me through ups and downs of my life in IUCAA, and<br />

patiently tolerated years of living far apart. She listened to my ideas carefully, asked<br />

very difficult questions which motivated me to learn more, helped me in preparation of<br />

this thesis and designed fancy headers for my synopsis. I am thankful to her for her<br />

support and faith she showed in me and for the great patience she showed in the difficult<br />

times of our life.<br />

xviii


Contents<br />

Abstract<br />

Acknowledgements<br />

List of Figures<br />

List of Tables<br />

ix<br />

xiii<br />

xxi<br />

xxiii<br />

1 Introduction 1<br />

2 Statistical Properties of the <strong>CMB</strong> <strong>Anisotropy</strong> 6<br />

2.1 Statistical Isotropy of Temperature Anisotropies . . . . . . . . . . . . . . 8<br />

2.2 Statistical Isotropy and Gaussianity . . . . . . . . . . . . . . . . . . . . . 10<br />

2.3 The Bipolar Power Spectrum (BiPS) . . . . . . . . . . . . . . . . . . . . 11<br />

2.4 BiPS Representation in Real Space . . . . . . . . . . . . . . . . . . . . . 13<br />

2.5 Unbiased Estimator of BiPS . . . . . . . . . . . . . . . . . . . . . . . . . 16<br />

2.6 Cosmic Variance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18<br />

2.7 Sources of Breakdown of Statistical Isotropy . . . . . . . . . . . . . . . . 20<br />

2.8 Primordial Magnetic Fields . . . . . . . . . . . . . . . . . . . . . . . . . 21<br />

3 SI Violation: Analytic Models 23<br />

3.1 One Preferred Axis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23<br />

3.2 Three Preferred Axes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26<br />

3.3 Oblique Preferred Axis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27<br />

3.4 Toy Model With Cross Terms . . . . . . . . . . . . . . . . . . . . . . . . 29<br />

3.5 Fourth Order Toy Models With Three Axes . . . . . . . . . . . . . . . . 30<br />

3.6 Toy Model II . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31<br />

3.7 Toroidal Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32<br />

4 Signatures of Cosmic Topology 33<br />

4.1 Statistical Isotropy Violation in Compact Spaces . . . . . . . . . . . . . . 34<br />

xix


4.2 Method of images . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39<br />

4.3 A Bipolar Power Spectrum Study of Cosmic Topology . . . . . . . . . . . 41<br />

4.3.1 Flat Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42<br />

4.3.2 Spherical Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . 63<br />

5 Homogeneous Cosmological Models 75<br />

5.1 Classification of Cosmological Models . . . . . . . . . . . . . . . . . . . . 77<br />

5.2 Modern Classification of Bianchi Models . . . . . . . . . . . . . . . . . . 80<br />

5.3 The Field Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81<br />

5.4 Approaching Isotropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84<br />

5.5 Energy Momentum Tensor and Vorticity Vector . . . . . . . . . . . . . . 85<br />

5.6 Null Geodesics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87<br />

5.7 Patterns of Anisotropic Models on the <strong>CMB</strong> Sky . . . . . . . . . . . . . . 89<br />

5.7.1 <strong>CMB</strong> Patterns in Type I Bianchi Models . . . . . . . . . . . . . . 91<br />

5.7.2 <strong>CMB</strong> Patterns in Type V Bianchi Models . . . . . . . . . . . . . 91<br />

5.7.3 <strong>CMB</strong> Patterns in Type VII 0 Bianchi Models . . . . . . . . . . . . 92<br />

5.7.4 <strong>CMB</strong> Patterns in Type VII h Bianchi Models . . . . . . . . . . . . 92<br />

5.7.5 <strong>CMB</strong> Patterns in Type IX Bianchi Models . . . . . . . . . . . . . 95<br />

5.8 Unveiling Hidden Patterns of Bianchi VII h . . . . . . . . . . . . . . . . . 96<br />

6 Observational Artifacts 105<br />

6.1 Anisotropic noise . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105<br />

6.2 The effect of non-circular beam . . . . . . . . . . . . . . . . . . . . . . . 108<br />

6.3 Mask effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109<br />

6.4 Residuals from foreground removal . . . . . . . . . . . . . . . . . . . . . 110<br />

7 Statistical isotropy of <strong>CMB</strong> anisotropy from WMAP 111<br />

8 Summary and future directions 127<br />

A Detailed Steps of Calculations 131<br />

B Cosmic Variance of BiPS, Handling Gaussian 8 Point Functions 139<br />

C Search for Planar Symmetries in <strong>CMB</strong> Ansiotropy Maps 148<br />

D Symmetries of the Space 153<br />

E Bianchi Models: Bianchi’s Original Classification 158<br />

F Simulating <strong>CMB</strong> maps for non-SI correlation functions 165<br />

G Useful Standard Integrals and Summation Rules 172<br />

Bibliography 174<br />

xx


List of Figures<br />

2.1 BipoSH coefficients . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12<br />

2.2 <strong>CMB</strong> temperature correlation function . . . . . . . . . . . . . . . . . . . 14<br />

2.3 Comparing analytically calculated bias <strong>with</strong> numerical computations . . . 17<br />

2.4 Analytically calculated cosmic variance vs. numerical computations . . . 19<br />

2.5 Effect of magnetic fields on covariance matrix . . . . . . . . . . . . . . . 21<br />

2.6 BiPS of homogeneous magnetic fields . . . . . . . . . . . . . . . . . . . . 22<br />

4.1 Fundamental domains for compact flat spaces . . . . . . . . . . . . . . . 44<br />

4.2 The chimney space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45<br />

4.3 The slab space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45<br />

4.4 Two point correlation in T 2 . . . . . . . . . . . . . . . . . . . . . . . . . 47<br />

4.5 Folded circle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48<br />

4.6 Correlations along folded circle . . . . . . . . . . . . . . . . . . . . . . . 48<br />

4.7 Circles in the sky . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49<br />

4.8 Dirichlet domains of squeezed tori . . . . . . . . . . . . . . . . . . . . . . 50<br />

4.9 BiPS of toroidal spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51<br />

4.10 Geometry of the distance between two points . . . . . . . . . . . . . . . . 53<br />

4.11 <strong>CMB</strong> temperature correlation function. . . . . . . . . . . . . . . . . . . . 64<br />

4.12 Fundamental domain for the regular dodecahedron . . . . . . . . . . . . 68<br />

4.13 Two point correlations in Poincaré dodecahedral spaces . . . . . . . . . . 69<br />

4.14 BiPS of Poincaré dodecahedron . . . . . . . . . . . . . . . . . . . . . . . 72<br />

4.15 Simulated <strong>CMB</strong> maps of Poincaré dodecahedral spaces . . . . . . . . . . 73<br />

4.16 BiPS of simulated <strong>CMB</strong> maps of Poincaré dodecahedral spaces . . . . . . 74<br />

5.1 BiPS of Bianchi VII h templates . . . . . . . . . . . . . . . . . . . . . . . 78<br />

5.2 Bianchi VII h patterns for x = 0.1 . . . . . . . . . . . . . . . . . . . . . . 96<br />

5.3 Bianchi VII h patterns for x = 0.55 . . . . . . . . . . . . . . . . . . . . . . 98<br />

5.4 Bianchi VII h patterns for x = 1 . . . . . . . . . . . . . . . . . . . . . . . 98<br />

5.5 Bianchi VII h patterns for x = 5 . . . . . . . . . . . . . . . . . . . . . . . 98<br />

5.6 Bianchi corrected map . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99<br />

5.7 Comparison of power spectra . . . . . . . . . . . . . . . . . . . . . . . . 99<br />

xxi


5.8 Adding a template to the background map . . . . . . . . . . . . . . . . . 100<br />

5.9 Bianchi added maps, Ω 0 = 0.5 . . . . . . . . . . . . . . . . . . . . . . . . 101<br />

5.10 BiPS of Bianchi added maps, Ω 0 = 0.5 . . . . . . . . . . . . . . . . . . . 102<br />

5.11 Probability distribution of χ 2 for low density Bianchi template . . . . . . 103<br />

5.12 Bianchi added maps, Ω 0 = 0.99 . . . . . . . . . . . . . . . . . . . . . . . 103<br />

5.13 BiPS of Bianchi added maps, Ω 0 = 0.99 . . . . . . . . . . . . . . . . . . . 104<br />

5.14 Probability distribution of χ 2 for nearly flat Bianchi template . . . . . . 104<br />

6.1 Coadded noise covariance matrix of WMAP . . . . . . . . . . . . . . . . 106<br />

6.2 BiPS of diagonal noise covariance matrix . . . . . . . . . . . . . . . . . . 107<br />

6.3 Estimating BiPS from simulations of noise maps . . . . . . . . . . . . . . 108<br />

7.1 Window functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116<br />

7.2 BiPS of WMAP . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117<br />

7.3 BiPS of WMAP . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118<br />

7.4 Effect of Galactic mask . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119<br />

7.5 Probability distribution function of BiPS . . . . . . . . . . . . . . . . . . 120<br />

7.6 Probability of <strong>CMB</strong> maps being SI . . . . . . . . . . . . . . . . . . . . . 121<br />

7.7 SI violation vanishes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121<br />

7.8 Probability of <strong>CMB</strong> maps being SI . . . . . . . . . . . . . . . . . . . . . 122<br />

7.9 BiPS of WMAP filtered <strong>with</strong> bandpass filters . . . . . . . . . . . . . . . 123<br />

7.10 BiPS of WMAP filtered <strong>with</strong> Gaussian filters . . . . . . . . . . . . . . . . 124<br />

7.11 PDF of κ l for a Gaussian filter Wl G (40) . . . . . . . . . . . . . . . . . . . 125<br />

7.12 PDF of κ l for a band pass filter Wl S (20, 30) . . . . . . . . . . . . . . . . . 126<br />

C.1 S-MAP . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151<br />

C.2 S 0 statistics for ILC map . . . . . . . . . . . . . . . . . . . . . . . . . . . 152<br />

F.1 Simulated noise map . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168<br />

F.2 Simulated map of <strong>CMB</strong> temperature anisotropy in a toroidal universe . . 171<br />

xxii


List of Tables<br />

4.1 Classification of three-dimensional flat spaces . . . . . . . . . . . . . . . . 43<br />

4.2 Finite subgroups of Q 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66<br />

5.1 Classification of Cosmological Models <strong>with</strong> Respect to Their Symmetries. 79<br />

5.2 Classification of Homogeneous Cosmological Models into Ten Equivalence<br />

Classes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81<br />

xxiii


Chapter 1<br />

Introduction<br />

This thesis is a study of statistics of the <strong>CMB</strong> anisotropy as a Gaussian random field<br />

on a sphere. We study statistical isotropy (SI) of <strong>CMB</strong> temperature anisotropy and use<br />

that to draw interesting conclusions on the physics, topology and large scale structure<br />

of the universe as well as properties of the systematics in <strong>CMB</strong> experiments.<br />

The Cosmic Microwave Background (<strong>CMB</strong>) anisotropy is proved to be a very powerful<br />

observational probe of cosmology. Detailed measurements of the anisotropies in<br />

the <strong>CMB</strong> can provide a wealth of information about the global properties, constituents,<br />

history and eventual fate of the Universe. In standard cosmology, the <strong>CMB</strong> anisotropy<br />

is expected to be statistically isotropic, i.e., statistical expectation values of the temperature<br />

fluctuations ∆T (ˆq) are preserved under rotations of the sky. In particular,<br />

the angular correlation function C(ˆq, ˆq ′ ) ≡ 〈∆T (ˆq)∆T (ˆq ′ )〉 is rotationally invariant for<br />

Gaussian fields. In spherical harmonic space, where ∆T (ˆq) = ∑ lm a lmY lm (ˆq) this translates<br />

to a diagonal 〈a lm a ∗ l ′ m ′〉 = C lδ ll ′δ mm ′ where C l is the widely used angular power<br />

spectrum of <strong>CMB</strong> anisotropy. Statistical isotropy (SI) is usually assumed in almost all<br />

<strong>CMB</strong> studies. But there are certain mechanisms that violate this condition. This thesis<br />

is dedicated to the study of this fundamental assumption in <strong>CMB</strong>. We have formulated<br />

the problem mathematically, constructed a quantitative method to describe SI violation,<br />

studied possible ways of violation of SI condition and finally examined this fundamental<br />

assumption using the most recent <strong>CMB</strong> anisotropy data.<br />

In February 2003, Wilkinson Microwave <strong>Anisotropy</strong> Probe (WMAP) released its first<br />

data (Bennett et al. , 2003; Hinshaw et al. , 2003; Kogut et al. , 2003; Page et al. , 2003;<br />

1


1.0 2<br />

Peiris et al. , 2003; Spergel et al. , 2003). The WMAP mission was designed to advance<br />

observational cosmology by making accurate full sky <strong>CMB</strong> maps <strong>with</strong> precision and<br />

reliability. WMAP heralds the dawn of precision cosmology tightly constraining many<br />

key parameters (Bridle et al., 2003). Determining the flatness of space, the near scaleinvariance,<br />

adiabaticity, and Gaussian distribution of the density perturbations, the<br />

density of baryons, the age of the universe, and the early formation of the first stars are<br />

some major results of WMAP and are consistent <strong>with</strong> the standard model of cosmology<br />

(for an interesting review on standard model see Scott (2005)). After the release of firstyear<br />

data of WMAP, statistical isotropy of the <strong>CMB</strong> anisotropy attracted considerable<br />

attention. Tantalizing evidence of SI breakdown (albeit, in very different guises) has<br />

mounted in the WMAP first year sky maps, using a variety of different statistics. It was<br />

pointed out that the suppression of power in the quadrupole and octopole are aligned<br />

(Tegmark et al. , 2004). Further “multipole-vector” directions associated <strong>with</strong> these<br />

multipoles (and some other low multipoles as well) appear to be anomalously correlated<br />

(Copi et al. , 2004; Schwarz et al. , 2004). There are indications of asymmetry in<br />

the power spectrum at low multipoles in opposite hemispheres (Eriksen et al. , 2004a;<br />

Hansen et al. , 2004; Naselsky et al. , 2004). Possibly related, are the results of tests<br />

of Gaussianity that show asymmetry in the amplitude of the measured genus amplitude<br />

(at about 2 to 3σ significance) between the north and south galactic hemispheres (Park<br />

, 2004; Eriksen et al. , 2004b,c). Analysis of the distribution of extrema in WMAP<br />

sky maps has indicated non-Gaussianity, and to some extent, violation of SI (Larson &<br />

Wandelt , 2004). However, what is missing is a common, well defined, mathematical<br />

language to quantify SI (as distinct from non Gaussianity) and the ability to ascribe<br />

statistical significance to the anomalies unambiguously.<br />

In order to extract more information from the rich source of information provided<br />

by present (and future) <strong>CMB</strong> maps, it is important to design as many independent<br />

statistical methods as possible to study deviations from standard statistics, such as<br />

statistical isotropy. Since SI can be violated in many different ways, various statistical<br />

methods can come to different conclusions. Because each method by design is more<br />

sensitive to a special kind of SI violation.<br />

We have constructed a novel statistical tool of searching for departures from SI.


1.0 3<br />

This method is now known as the bipolar power spectrum (BiPS) which is sensitive<br />

to structures and patterns in the underlying two-point correlation function (Hajian &<br />

Souradeep , 2005, 2003b; Souradeep & Hajian , 2003). The BiPS is particularly sensitive<br />

to real space correlation patterns (preferred directions, etc.) on characteristic angular<br />

scales. In harmonic space, the BiPS at multipole l sums power in off-diagonal elements<br />

of the covariance matrix, 〈a lm a l ′ m ′〉, in the same way that the ‘angular momentum’<br />

addition of states lm, l ′ m ′ have non-zero overlap <strong>with</strong> a state <strong>with</strong> angular momentum<br />

|l−l ′ | < l < l+l ′ . Signatures, like a lm and a l+nm being correlated over a significant range<br />

l are ideal targets for BiPS. These are typical of SI violation due to cosmic topology<br />

and the predicted BiPS in these models have a strong spectral signature in the bipolar<br />

multipole l space (Hajian & Souradeep , 2003a). The orientation independence of BiPS<br />

is an advantage since one can constrain mechanisms that violate the SI independent of<br />

the unknown specific orientation of the pattern (e.g., preferred directions). The BiPS<br />

statistics has been applied to the most recent full sky maps of the <strong>CMB</strong> anisotropy.<br />

The null detection allows us to place observational constraints on exotic possibilities for<br />

cosmology such as compact spaces, anisotropic cosmological models, etc.<br />

Here I give a brief description of how the thesis is organized into chapters.<br />

Chapter 2 contains the basic material and is provided as a qualitative introduction<br />

to the statistical isotropy. In this chapter we introduce basic statistical properties of<br />

<strong>CMB</strong> and discuss the statistical isotropy as distinct from Gaussianity. Also we introduce<br />

BiPS in real as well as harmonic space to quantify statistical isotropy violations in a<br />

<strong>CMB</strong> map. We define an unbiased estimator for BiPS and compute its cosmic variance<br />

in details. We introduce homogeneous primordial magnetic fields as the first example of<br />

a theoretical model that does not respect the SI and study the signature of these models<br />

on the bipolar power spectrum. A homogeneous primordial magnetic field <strong>with</strong> fixed<br />

direction induces correlations between the a l−1,m and a l+1,m multipole coefficients of the<br />

<strong>CMB</strong> temperature anisotropy field (see e.g. Durrer et al.<br />

(1998)). This means that a<br />

homogeneous primordial magnetic field <strong>with</strong> fixed direction breaks down the statistical<br />

isotropy in its own manner and hence its bipolar power spectrum has a specific shape.<br />

In Chapter 3 we study toy models of non-SI correlations which are analytically<br />

solvable. These toy models are motivated by various mechanisms of SI violation such as


1.0 4<br />

multiconnected spaces. Besides helping us to come to a good understanding of violation<br />

of SI, they shed a light on how BiPS method works. We repeatedly refer to this chapter<br />

to explain several signatures of different models on BiPS.<br />

In Chapters 4 to 6 we study possible sources of breakdown of SI. These sources<br />

vary from pure theoretical models of ultra large scale structure of the universe to the<br />

systematics of <strong>CMB</strong> experiments.<br />

Chapter 4 is dedicated to the study of topology of the universe. The cleanest signature<br />

of the topology of the universe is written on the microwave sky. We use BiPS to<br />

look for these signatures in a <strong>CMB</strong> map. BiPS exploits the existence of preferred directions<br />

along the highly correlated spots in compact spaces. Number and configuration<br />

of these preferred directions differ in each compact space and cause different correlation<br />

patterns on the <strong>CMB</strong> sky. We have shown that bipolar power spectrum is very<br />

sensitive to these patterns and can be used to distinguish them. A detailed study of<br />

the flat spaces has been carried out in this chapter. It is shown that different multiconnected<br />

spaces have different BiPS signatures. These signatures are found analytically<br />

and verified by numerical computation of BiPS on multiply connected flat spaces. In<br />

the second half of this chapter we study Poincare Dodecahedral spaces suggested as a<br />

possible explanation of low quadrupole of the WMAP data by Luminet et al. (2003).<br />

We show that the Poincaré Dodecahedron <strong>with</strong> Ω k = 0.013 breaks down the statistical<br />

isotropy at an observable level which was not observed in the WMAP data. Our results<br />

are consistent <strong>with</strong> the complete Bayesian likelihood analysis of Poincaré Dodecahedral<br />

spaces [Hajian et al. (2006)].<br />

Anisotropic universe is another example of violation of SI. In Chapter 5 we study<br />

homogeneous cosmological models which are initially anisotropic and as the time goes<br />

on they become more isotropic and will asymptotically tend to FRW models. Bianchi<br />

models provide a generic description of homogeneous anisotropic cosmologies, see e.g.<br />

Ellis & MacCallum (1969) and Collins & Hawking (1973b). Distinct features of Bianchi<br />

models studied by a number of authors (Barrow et al. , 1985; Bunn et al. , 1996; Kogut et<br />

al. , 1997) and were constrained. Recently they were revived after the first year WMAP<br />

data by Jaffe et al. (2005). The authors reported the discovery of a hidden pattern<br />

in the WMAP data which according to them was due to a rotating universe template.


1.0 5<br />

The model studied by Jaffe et al. (2005) was not self-consistent as a rotating universe<br />

model. But the existence of a hidden template <strong>with</strong> a preferred axis in the WMAP data<br />

is very well suited the BiPS method. We have shown that hiding a Bianchi template in<br />

a ΛCDM <strong>CMB</strong> anisotropy temperature map, will lead to measurable non-zero bipolar<br />

power spectra. The null BiPS of WMAP can be used to put tight constraints on the<br />

existence of such hidden patterns in the WMAP data.<br />

In addition to the theoretical models presented above, foregrounds and observational<br />

artifacts (such as non-circular beam, incomplete/non-uniform sky coverage and<br />

anisotropic noise) would also manifest themselves as violations of SI. Chapter 6 is<br />

devoted to the study of these observational effects.<br />

In Chapter 7 we discuss the results of applying the BiPS method to the first year<br />

WMAP data. We carried out measurement of the BiPS, on four full-sky <strong>CMB</strong> anisotropy<br />

maps. Using various window functions we isolated regions in multipole space and probed<br />

them independently. Our results indicate that there is no strong evidence for violation<br />

of statistical isotropy in the first-year WMAP data on angular scales larger than that<br />

corresponding to l ∼ 60. We have verified that our null results are consistent <strong>with</strong><br />

simulated statistically isotropic maps (1000 realizations). These results also confirm the<br />

other observational artifacts that cause SI violation in the <strong>CMB</strong> maps of WMAP, such<br />

as, incomplete and non-uniform sky coverage, beam anisotropy, foreground residuals and<br />

statistically anisotropic noise do not significantly contribute to the map (up to l ∼ 60).<br />

In the end we have briefly discussed the use of BiPS in <strong>CMB</strong> polarization studies.<br />

In order to make this thesis self contained, we have included details of calculations,<br />

mathematical preliminaries and computational methods. However, these are given in<br />

the appendices to keep the thesis readable. In addition to the above appendices, we<br />

present two methods of generating <strong>CMB</strong> sky maps from anisotropic correlation functions<br />

which have been used throughout this work in Appendix F. These methods are<br />

diagonalization method and Cholesky decomposition method. In these methods we use<br />

the full covariance matrix to generate the <strong>CMB</strong> map. This is very important when we<br />

deal <strong>with</strong> statistically anisotropic models such as those discussed above. We also present<br />

a pedagogical map making recipe for the beginners to make their own maps very easily.


Chapter 2<br />

Statistical Properties of the <strong>CMB</strong><br />

<strong>Anisotropy</strong><br />

The temperature anisotropies of the <strong>CMB</strong> are described by a random field, ∆T (ˆn) =<br />

T (ˆn) − T 0 , on a 2-dimensional surface of a sphere (the sky), where ˆn = (θ, φ) is a unit<br />

vector on the sphere and T 0 = ∫ dΩˆn<br />

T (ˆn) represents the mean temperature of the <strong>CMB</strong>.<br />

4π<br />

The statistical properties of this field can be characterized by the n-point correlation<br />

functions<br />

〈∆T (ˆn 1 )∆T (ˆn 2 ) · · · ∆T (ˆn n )〉. (2.1)<br />

Here the bracket denotes the ensemble average, i.e. an average over all possible configurations<br />

of the field. If the field is Gaussian in nature, the connected part of the n-point<br />

functions disappears for n > 2. The non-zero (even-n)-point correlation functions can<br />

be expressed in terms of the 2-point correlation function. As a result, a Gaussian distribution<br />

is completely described by the two-point correlation function.<br />

The statistics of the temperature fluctuations is Gaussian because they are linearly<br />

connected to quantum vacuum fluctuations of a very weakly interacting scalar field (the<br />

inflaton). Perturbations in primordial gravitational potential, Φ 0 , generate the <strong>CMB</strong><br />

anisotropy, ∆T . The linear perturbation theory gives a linear relation between Φ 0<br />

and ∆T . Φ 0 (x) is assumed to be Gaussian, because in inflation, quantum fluctuations<br />

produce Gaussian random fluctuations in the gravitational potential<br />

∫ d 3 k<br />

Φ 0 (x) =<br />

(2π) Φ 0(k) exp ik · x (2.2)<br />

3<br />

6


2.0 7<br />

<strong>with</strong> 〈Φ 0 (k)Φ ∗ 0 (k′ )〉 = P (k)δ(k − k ′ ), the power spectrum of potential fluctuations at<br />

that epoch.<br />

In the linear approximation in fluctuations, in a spatially flat universe,<br />

the temperature anisotropies at the surface of last scattering are linearly related to the<br />

potential fluctuations at that epoch 1<br />

∫<br />

∆T (ˆn) =<br />

d 3 k<br />

(2π) 3 eik·ˆnτrec Φ 0 (k)g T (k), (2.3)<br />

where τ rec is the comoving conformal lookback time of the recombination epoch (τ is the<br />

conformal time τ = ∫ dt/a), and g T (k) is the linear radiation transfer function which<br />

describes the relationship between potential fluctuations and temperature fluctuations<br />

at recombination.<br />

For small k, g T (k) = 1/3, which is the well known fact that for<br />

temperature fluctuations on super-horizon scales at the decoupling epoch, the Sachs-<br />

Wolfe effect (Sachs and Wolfe , 1967) dominates and hence on large angular scales<br />

On sub-horizon scales, g T<br />

∆T<br />

T (ˆn) = 1 3 Φ LS(ˆnτ rec , τ 0 ) (2.4)<br />

oscillates (due to acoustic oscillations in the baryon-photon<br />

fluid), and we need to evolve the coupled matter-radiation fluid, i.e. solve the Boltzmann<br />

photon transport equations coupled <strong>with</strong> the Einstein equations for g T .<br />

<strong>CMB</strong>FAST 2 compute g T (k) for different cosmological models.<br />

If we expand the gravitational potential into a series in powers of a density enhancement<br />

Codes like<br />

Φ(x, t) = Φ (1) + Φ (2) + · · · (2.5)<br />

The linear term Φ 1 = Φ 0 (x) and is assumed to be Gaussian. Non-linearity in inflation<br />

makes Φ weakly non-Gaussian and will lead to non-linear terms, such as Φ (2) , in eqn.<br />

(2.5). However, since potential fluctuations in the early universe are small, 〈Φ 2 〉 ∼ 10 −9 ,<br />

these non-linear effects are usually viewed as undetectable.<br />

Moreover starting from<br />

Gaussian initial condition, it has been shown that non-linear evolution does not produce<br />

seeds of non Gaussianity (Munshi et al. , 1995; Spergel et al. , 1999) (for a nice review<br />

on non-Gaussianity see Bartolo et al. (2004)). Hence, here we limit our study to<br />

1 This description is very simple but naive. It does not take the line of sight effects such as integrated<br />

Sachs-Wolfe effect into account. For a more careful description see Appendix A.<br />

2 Available at http://www.cmbfast.org/


2.1 STATISTICAL ISOTROPY OF TEMPERATURE ANISOTROPIES 8<br />

Gaussian <strong>CMB</strong> anisotropy field, where the two-point correlation function<br />

C(ˆn, ˆn ′ ) ≡ 〈∆T (ˆn)∆T (ˆn ′ )〉 (2.6)<br />

contains all the statistical information encoded in the field.<br />

It is convenient to expand the temperature anisotropy field into spherical harmonics,<br />

the orthonormal basis on the sphere, as<br />

∆T (ˆn) = ∑ l,m<br />

a lm Y lm (ˆn) , (2.7)<br />

where the complex quantities, a lm are given by<br />

∫<br />

a lm = dΩˆn Ylm ∗ (ˆn)∆T (ˆn). (2.8)<br />

For a Gaussian <strong>CMB</strong> anisotropy, a lm are Gaussian random variables and hence, the<br />

covariance matrix, 〈a lm a ∗ l ′ m ′〉, fully describes the whole field.<br />

2.1 Statistical Isotropy of Temperature Anisotropies<br />

A random field Φ(r) is statistically homogeneous (in the wide sense) if the mean<br />

and the second moment (covariance) of that remain invariant under the coordinate<br />

transformation r → r + δr. That is to say,<br />

〈Φ(r)〉 = 〈Φ(r + δr)〉, (2.9)<br />

C Φ (r 1 , r 2 ) = C Φ (r 1 + δr, r 2 + δr).<br />

The first condition implies that 〈Φ〉 = const. Putting δr = −r 2 in the second condition<br />

we find that C Φ (r 1 , r 2 ) = C Φ (r 1 − r 2 , 0), i.e. the covariance of a statistically<br />

homogeneous field is dependent only on the difference r 1 − r 2 , and not on r 1 and r 2<br />

separately. So instead of C Φ (r 1 − r 2 , 0) we can simply write C Φ (r) which means:<br />

C Φ (r) = 〈˜Φ(r 1 )˜Φ(r 1 + r)〉, (2.10)<br />

where ˜Φ ≡ Φ − 〈Φ〉.<br />

Statistically homogeneous fields, <strong>with</strong> C Φ (r) dependent only on the magnitude(but<br />

not on the direction) of the vector r = r 2 − r 1 connecting the points r 1 and r 2<br />

C Φ (r) = C Φ (r), r = |r| = √ r · r, (2.11)


2.1 STATISTICAL ISOTROPY OF TEMPERATURE ANISOTROPIES 9<br />

are said to be statistically isotropic. The covariance of isotropic random fields is even<br />

and always real.<br />

For temperature anisotropies of the <strong>CMB</strong>, statistical isotropy is simply equivalent<br />

to rotational invariance of n-point correlation functions<br />

〈∆T (Rˆn 1 )∆T (Rˆn 2 ) · · · ∆T (Rˆn n )〉 = 〈∆T (ˆn 1 )∆T (ˆn 2 ) · · · ∆T (ˆn n )〉, (2.12)<br />

where ∆T (Rˆn 1 ) is the temperature anisotropy at Rˆn 1 which is pixel ˆn 1 rotated by the<br />

rotation matrix R(α, β, γ) for the Euler angles α, β and γ. For two point correlations<br />

of <strong>CMB</strong> anisotropy, the condition of statistical isotropy (eqn. 2.11) will translate into<br />

C(ˆn, ˆn ′ ) = C(θ), θ = arccos (ˆn · ˆn ′ ) (2.13)<br />

which implies that the two point correlation is just a function of separation angle, θ,<br />

between the two points on the sky.<br />

Legendre polynomials,<br />

C(θ) = 1<br />

4π<br />

It is then convenient to expand it in terms of<br />

∞∑<br />

(2l + 1)C l P l (cos θ), (2.14)<br />

l=2<br />

where C l is the widely used angular power spectrum. The summation over l starts from<br />

2 because the l = 0 term is the monopole which in the case of statistical isotropy the<br />

monopole is constant, and it can be subtracted out. The dipole l = 1 is due to the<br />

local motion of the observer and is subtracted out as well.<br />

To obtain the statistical isotropy condition in harmonic space, we expand the left<br />

hand side of eqn. (2.12) in terms of spherical harmonics, each ∆T (Rˆn i ) may be expanded<br />

as<br />

∆T (Rˆn 1 ) = ∑ l,m<br />

= ∑ l,m<br />

a R lm Y lm(Rˆn 1 ) (2.15)<br />

a R lm<br />

m ′<br />

∑<br />

Y lm ′(ˆn 1 )Dm l m(R),<br />

′<br />

in which a R lm are harmonic coefficients in rotated coordinates and Dl m ′ m (R) is the finite<br />

rotation matrix in the lm-representation, Wigner D-function Varshalovich et al. (1988).<br />

By substituting this into eqn.(2.12) and using the orthonormality law of spherical har-


2.2 STATISTICAL ISOTROPY AND GAUSSIANITY 10<br />

monics, eqn.(G.2), we obtain the statistical isotropy condition in harmonic space<br />

∑<br />

m ′ 1 ,m′ 2···,m′ n<br />

〈a R l 1 m ′ aR 1 l 2 m · · · ′ aR 2 l nm ′ 〉Dl 1<br />

n m 1 m<br />

(R)D l ′ 2<br />

1 m 2 m<br />

· · · D<br />

2(R) ln ′ m (R) = 〈a nm ′ l<br />

n 1 m 1<br />

a l2 m 2<br />

· · · a lnm n<br />

〉.<br />

(2.16)<br />

For a Gaussian field, only the covariance, 〈a lm a ∗ l ′ m ′〉, is important. As it is shown in<br />

Appendix A, the condition for statistical isotropy translates to a diagonal covariance<br />

matrix,<br />

〈a lm a ∗ l ′ m ′〉 = C lδ ll ′δ mm ′, (2.17)<br />

which means that the angular power spectrum of <strong>CMB</strong> anisotropy, C l , tells us everything<br />

we need to know about the (Gaussian) <strong>CMB</strong> anisotropy. When 〈a lm a ∗ l ′ m ′〉 is not diagonal,<br />

the angular power spectrum C l does not have all the information of the field and one<br />

should also take the off-diagonal elements into account.<br />

2.2 Statistical Isotropy and Gaussianity<br />

Statistical isotropy and Gaussianity of <strong>CMB</strong> anisotropies are two independent and<br />

distinct concepts. The Gaussianity of the primordial fluctuations is a key assumption<br />

of modern cosmology, motivated by simple models of inflation. And since the statistical<br />

properties of the primordial fluctuations are closely related to those of the cosmic microwave<br />

background (<strong>CMB</strong>) radiation anisotropy, a measurement of non-Gaussianity of<br />

the <strong>CMB</strong> will be a direct test of the inflation paradigm. If <strong>CMB</strong> anisotropy is Gaussian,<br />

then the two point correlation fully specifies the statistical properties. In the case of<br />

non-Gaussian fields, one must take the higher moments of the field into account in order<br />

to fully describe the whole field. Statistical isotropy is a condition which guarantees<br />

that statistical properties of <strong>CMB</strong> sky (e.g. n-point correlations) are invariant under<br />

rotation. Hence, in a SI model, the iso-contours of two point correlations where one<br />

point is fixed and the other scans the whole sky, are concentric circles around the given<br />

point. Any breakdown of statistical isotropy will make these iso-contours non-circular.<br />

In addition to cosmological mechanisms, instrumental and environmental effects in observations<br />

easily produce spurious breakdown of statistical isotropy.<br />

A Gaussian field may or may not respect the statistical isotropy. In other words we


2.3 THE BIPOLAR POWER SPECTRUM (BIPS) 11<br />

can have<br />

1. Gaussian and SI models: In these models two point correlation and C l have all<br />

the information of the field and C(ˆn 1 , ˆn 2 ) = C(ˆn 1 · ˆn 2 )<br />

2. Gaussian but not SI models: two point correlation contains all the information of<br />

the field, but C(ˆn 1 , ˆn 2 ) ≠ C(ˆn 1 · ˆn 2 ). So C l is not adequate to describe the field<br />

and off-diagonal elements of covariance matrix should be taken into account.<br />

3. Non-Gaussian but SI models: two point correlation alone is not sufficient to describe<br />

the whole field and we need to take higher moments into account but every<br />

n-point correlation of the field is invariant under rotation.<br />

4. Non-Gaussian and non-SI models: neither two point correlation has all the information<br />

nor it is invariant under rotation.<br />

Here we confine ourselves to Gaussian fields where we only need to consider two point<br />

correlations. We will present a general way of describing fields of the kind (1) and (2)<br />

in the above list.<br />

2.3 The Bipolar Power Spectrum (BiPS)<br />

Two point correlation of <strong>CMB</strong> anisotropies, C(ˆn 1 , ˆn 2 ), is a two point function on<br />

S 2 × S 2 , and hence can be expanded as<br />

C(ˆn 1 , ˆn 2 ) =<br />

∑<br />

l 1 ,l 2 ,L,M<br />

A lM<br />

l 1 l 2<br />

{Y l1 (ˆn 1 ) ⊗ Y l2 (ˆn 2 )} lM , (2.18)<br />

where A lM<br />

l 1 l 2<br />

are coefficients of the expansion (hereafter BipoSH coefficients) and {Y l1 (ˆn 1 )⊗<br />

Y l2 (ˆn 2 )} lM are the bipolar spherical harmonics which transform in the same manner as<br />

the spherical harmonic function <strong>with</strong> l, M <strong>with</strong> respect to rotations (Varshalovich et al.<br />

, 1988) given by<br />

{Y l1 (ˆn 1 ) ⊗ Y l2 (ˆn 2 )} lM = ∑ m 1 m 2<br />

C lM<br />

l 1 m 1 l 2 m 2<br />

Y l1 m 1<br />

(ˆn 2 )Y l2 m 2<br />

(ˆn 2 ), (2.19)<br />

in which C lM<br />

l 1 m 1 l 2 m 2<br />

are Clebsch-Gordan coefficients. We can inverse-transform C(ˆn 1 , ˆn 2 )<br />

to get the A lM<br />

l 1 l 2<br />

by multiplying both sides of eqn.(2.18) by {Y l ′<br />

1<br />

(ˆn 1 ) ⊗ Y l ′<br />

2<br />

(ˆn 2 )} ∗ l ′ M ′ and


2.3 THE BIPOLAR POWER SPECTRUM (BIPS) 12<br />

integrating over all angles. Then the orthonormality of bipolar harmonics, eqn. (G.2),<br />

implies that<br />

A lM<br />

l 1 l 2<br />

=<br />

∫<br />

dΩˆn1<br />

∫<br />

dΩˆn2 C(ˆn 1 , ˆn 2 ) {Y l1 (ˆn 1 ) ⊗ Y l2 (ˆn 2 )} ∗ lM. (2.20)<br />

The above expression and the fact that C(ˆn 1 , ˆn 2 ) is symmetric under the exchange of<br />

ˆn 1 and ˆn 2 lead to the following symmetries of A lM<br />

l 1 l 2<br />

A lM<br />

l 2 l 1<br />

= (−1) (l 1+l 2 −L) A lM<br />

l 1 l 2<br />

, (2.21)<br />

A lM<br />

ll = A lM<br />

ll δ l,2k−1 , k = 0, 1, 2, · · · .<br />

We show in Appendix A that the Bipolar Spherical Harmonic (BipoSH) coefficients,<br />

A lM<br />

l 1 l 2<br />

, are in fact linear combinations of off-diagonal elements of the covariance matrix,<br />

A lM<br />

l 1 l 2<br />

= ∑ m 1 m 2<br />

〈a l1 m 1<br />

a ∗ l 2 m 2<br />

〉(−1) m 2<br />

C lM<br />

l 1 m 1 l 2 −m 2<br />

. (2.22)<br />

This clearly shows that A lM<br />

l 1 l 2<br />

completely represent the information of the covariance<br />

120<br />

120<br />

100<br />

100<br />

l(l+1)+m+1<br />

80<br />

60<br />

40<br />

l(l+1)+m+1<br />

80<br />

60<br />

40<br />

20<br />

20<br />

0<br />

0 20 40 60 80 100 120<br />

l(l+1)+m+1<br />

0<br />

0 20 40 60 80 100 120<br />

l(l+1)+m+1<br />

Figure 2.1 BipoSH coefficients are linear combinations of elements of the covariance<br />

matrix.<br />

Here A 2M<br />

ll ′<br />

(left) and A4M ll ′<br />

(right) are plotted to show how BiPS covers the<br />

off-diagonal elements of the covariance matrix in harmonic space.<br />

matrix. Fig. 2.1 shows how A 2M<br />

l 1 l 2<br />

and A 4M<br />

l 1 l 2<br />

matrix. When SI holds, the covariance matrix is diagonal, hence<br />

combine the elements of the covariance<br />

A lM<br />

ll ′ = (−1)l C l (2l + 1) 1/2 δ ll ′ δ l0 δ M0 , (2.23)


2.4 BIPS REPRESENTATION IN REAL SPACE 13<br />

BipoSH expansion is the most general way of studying two point correlation functions<br />

of <strong>CMB</strong> anisotropy. The well known angular power spectrum, C l is in fact a subset of<br />

BipoSH coefficients.<br />

It is impossible to measure all A lM<br />

l 1 l 2<br />

A 00<br />

l 1 l 2<br />

= (−1) l 1√<br />

2l1 + 1 C l1 δ l1 l 2<br />

. (2.24)<br />

individually because of cosmic variance. Combining<br />

BipoSH coefficients into Bipolar Power Spectrum reduces the cosmic variance 3 .<br />

BiPS of <strong>CMB</strong> anisotropy is defined as a convenient contraction of the BipoSH coefficients<br />

κ l = ∑<br />

l,l ′ ,M<br />

|A lM<br />

ll ′ |2 ≥ 0. (2.25)<br />

The BiPS has interesting properties. It is orientation independent, i.e. invariant under<br />

rotations of the sky. For models in which statistical isotropy is valid, BipoSH coefficients<br />

are given by eqn (2.24), and therefore SI condition implies a null BiPS, i.e. κ l = 0 for<br />

every l > 0,<br />

κ l = κ 0 δ l0 . (2.26)<br />

Non-zero components of BiPS imply the break down of statistical isotropy, and this<br />

introduces BiPS as a measure of statistical isotropy,<br />

STATISTICAL ISOTROPY =⇒ κ l = 0 ∀l ≠ 0. (2.27)<br />

2.4 BiPS Representation in Real Space<br />

Two point correlation of the <strong>CMB</strong> anisotropy is given by ensemble average over<br />

many universes (eqn.(2.6)). But in reality, there is only one universe and the ensemble<br />

average is meaningless unless the <strong>CMB</strong> sky is SI, where the two point correlation function<br />

C(θ) can be well estimated by the average product of temperature fluctuations in two<br />

directions ˆn and ˆn ′ whose angular separation is θ (see Fig. 2.2), this can be written as<br />

˜C(θ) =<br />

∫<br />

dΩˆn<br />

4π<br />

∫<br />

dΩˆn ′<br />

4π δ(ˆn · ˆn′ − cos θ)∆T (ˆn)∆T (ˆn ′ ) (2.28)<br />

3 This is similar to combining a lm to construct the angular power spectrum, C l = 1<br />

2l+1<br />

∑<br />

m |a lm| 2 ,<br />

to reduce the cosmic variance


2.4 BIPS REPRESENTATION IN REAL SPACE 14<br />

Figure 2.2 C(θ) ‘recovered’ from a map of a statistically isotropic model matches the<br />

original C(θ) shown as a solid line from which we made the map. The error bars shown<br />

are computed from 10 realizations. Note that the size of error bar depends on the size<br />

of θ bins.<br />

If the statistical isotropy is violated the estimate of the correlation function from a<br />

sky map given by a single temperature product<br />

˜C(ˆn 1 , ˆn 2 ) = ∆T (ˆn 1 )∆T (ˆn 2 ) (2.29)<br />

is poorly determined.<br />

Although it is not possible to estimate each element of the full correlation function<br />

C(ˆn 1 , ˆn 2 ), some measures of statistical isotropy of the <strong>CMB</strong> map can be estimated<br />

through suitably weighted angular averages of ∆T (ˆn 1 )∆T (ˆn 2 ). The angular averaging<br />

procedure should be such that the measure involves averaging over sufficient number of<br />

independent measurements , but should ensure that the averaging does not erase all the<br />

signature of statistical anisotropy. Another important desirable property is that measure


2.4 BIPS REPRESENTATION IN REAL SPACE 15<br />

be independent of the overall orientation of the sky. Based on these considerations, we<br />

have proposed a set of measures of statistical isotropy Hajian & Souradeep<br />

(2003b)<br />

which can be shown that is equivalent to the one in eqn.(2.25)<br />

∫ ∫ ∫<br />

κ l = (2l + 1) 2 1<br />

dΩ n1 dΩ n2 [ dRχ l (R) C(Rˆn<br />

8π 2 1 , Rˆn 2 )] 2 . (2.30)<br />

In the above expression, C(Rˆn 1 , Rˆn 2 ) is the two point correlation at Rˆn 1 and Rˆn 2<br />

which are the coordinates of the two pixels ˆn 1<br />

and ˆn 2 after rotating the coordinate<br />

system through an angle ω where (0 ≤ ω ≤ π) about the axis n(Θ, Φ). The direction of<br />

this rotation axis n is defined by the polar angles Θ where (0 ≤ Θ ≤ π) and Φ, where<br />

(0 ≤ Φ ≤ 2π). χ l is the trace of the finite rotation matrix in the lM-representation<br />

χ l (R) =<br />

l∑<br />

M=−l<br />

DMM l (R), (2.31)<br />

which is called the characteristic function, or the character of the irreducible representation<br />

of rank l. It is invariant under rotations of the coordinate systems. Explicit forms<br />

of χ l (R) are simple when R is specified by rotation by ω about an axis n(Θ, Φ), then<br />

χ l (R) is completely determined by the rotation angle ω and it is independent of the<br />

rotation axis n(Θ, Φ),<br />

χ l (R) = χ l (ω) (2.32)<br />

=<br />

sin [(2l + 1)ω/2]<br />

.<br />

sin [ω/2]<br />

And finally dR in eqn(2.30) is the volume element of the three-dimensional rotation<br />

group and is given by<br />

dR = 4 sin 2 ω dω sin Θ dΘ dΦ . (2.33)<br />

2<br />

As it is shown in Appendix (A), this expression for BiPS may be simplified further as<br />

κ l = 2l + 1 ∫ ∫<br />

∫<br />

dΩ<br />

8π 2 1 dΩ 2 C(ˆq 1 , ˆq 2 ) dRχ l (R)C(Rˆq 1 , Rˆq 2 ). (2.34)<br />

For a statistically isotropic model C(ˆn 1 , ˆn 2 ) is invariant under rotation, and therefore<br />

C(Rˆn 1 , Rˆn 2 ) = C(ˆn 1 , ˆn 2 ).<br />

If we use this property to substitute in eqn.(2.34) for<br />

C(Rˆn 1 , Rˆn 2 ), and if we use the orthonormality of χ l (ω) (see eqn. G.7), we will recover<br />

the condition for SI,<br />

κ l = κ 0 δ l0 . (2.35)


2.5 UNBIASED ESTIMATOR OF BIPS 16<br />

The proof of equivalence of BiPS definition in eqns.(2.30) and (2.25) is as follows. In<br />

eqn.(2.30), we substitute (2.31) for χ l (R) and from<br />

C(Rˆn 1 , Rˆn 2 ) =<br />

∑lM ′<br />

A lM ′<br />

l 1 l 2<br />

l 1 l 2<br />

∑<br />

DMM l ′(R){Y l 1<br />

(ˆn 1 ) ⊗ Y l2 (ˆn 2 )} lM , (2.36)<br />

M<br />

we substitute for C(Rˆn 1 , Rˆn 2 ) and use the orthonormality of DMM l (R) (see eqn.(G.8)),<br />

we will obtain eq.(2.25). Real-space representation of BiPS is very suitable for analytical<br />

computation of BiPS for theoretical models where we know the analytical expression for<br />

the two point correlation of the model, such as theoretical models in Hajian & Souradeep<br />

(2003a). This formalism would immensely simplify analytical calculations. On the other<br />

hand, the harmonic representation of BiPS allows computationally rapid methods for<br />

BiPS estimation from a given <strong>CMB</strong> map.<br />

2.5 Unbiased Estimator of BiPS<br />

In statistics, an estimator is a function of the known data that is used to estimate<br />

an unknown parameter; an estimate is the result from the actual application of the<br />

function to a particular set of data. Many different estimators may be possible for any<br />

given parameter. The above theory can be use to construct an estimator for measuring<br />

BipoSH coefficients from a given <strong>CMB</strong> map<br />

à lM<br />

ll ′ = ∑ mm ′ √<br />

Wl W l ′a lm a l ′ m ′ ClM lml ′ m ′ , (2.37)<br />

where W l is the Legendre transform of the window function. The above estimator is<br />

un-biased. Bias is the mismatch between ensemble average of the estimator and the<br />

true value. Bias for the BiPS is defined as B l = 〈˜κ l 〉 − κ l and is equal to<br />

B l = ∑ l 1 ,l 2<br />

W l1 W l2<br />

∑<br />

∑<br />

[<br />

〈a<br />

∗<br />

l1 m 1<br />

a l1 m ′ 1 〉〈a∗ l 2 m 2<br />

a l2 m ′ 2 〉 + 〈a∗ l 1 m 1<br />

a l2 m ′ 2 〉〈a∗ l 2 m 2<br />

a l1 m ′ 1 〉]<br />

m 1 ,m ′ 1 m 2 ,m ′ 2<br />

× ∑ M<br />

C lM<br />

l 1 m 1 l 2 m 2<br />

C lM<br />

l 1 m ′ 1 l 2m ′ 2 . (2.38)<br />

Therefore an un-biased estimator of BiPS is given by<br />

∣<br />

∣ ∣∣<br />

2 ∣ÃlM ll − ′ Bl , (2.39)<br />

˜κ l = ∑ ll ′ M


2.5 UNBIASED ESTIMATOR OF BIPS 17<br />

The above expression for B l is obtained by assuming Gaussian statistics of the temperature<br />

fluctuations. The procedure is very similar to computing cosmic variance (which<br />

is discussed in the next section), but much simpler. However, we can not measure the<br />

ensemble average in the above expression and as a result, elements of the covariance<br />

matrix (obtained from a single map) are poorly determined due to the cosmic variance.<br />

The best we can do is to compute the bias for the SI component of a map<br />

B l ≡ 〈˜κ B l 〉 SI<br />

= (2l + 1) ∑ ∑l+l 1<br />

l 1<br />

l 2 =|l−l 1 |<br />

C l1 C l2 W l1 W l2<br />

[<br />

1 + (−1) l δ l1 l 2<br />

]<br />

. (2.40)<br />

Note , the estimator ˜κ l is unbiased, only for SI correlation,i.e., 〈˜κ l 〉 = 0. Consequently,<br />

for SI correlation, the measured ˜κ l will be consistent <strong>with</strong> zero <strong>with</strong>in the error bars given<br />

by σ SI<br />

Hajian & Souradeep (2003b). We simulated 1000 SI <strong>CMB</strong> maps and computed<br />

BiPS for them using different filters. The average BiPS of SI maps is an estimation of<br />

the bias which can be compared to our analytical estimation. Fig. 2.3 shows that the<br />

theoretical bias (computed from average C l ) match the numerical estimations of average<br />

κ l of the 1000 realizations of the SI maps.<br />

Bias<br />

Average<br />

Bias<br />

Average<br />

0<br />

0<br />

5 10 15 20<br />

5 10 15 20<br />

Figure 2.3 Analytical bias for (left) a two beam window function <strong>with</strong> W S<br />

l (20, 30) and<br />

(right) a Gaussian window function <strong>with</strong> Wl<br />

G (40) computed from the average C l from<br />

1000 realizations of a SI <strong>CMB</strong> map compared <strong>with</strong> 〈κ realization<br />

l 〉 (the average κ l from<br />

1000 realizations). This shows that the theoretical bias is a very good estimation of the<br />

bias for a statistically isotropic map.<br />

It is important to note that bias cannot be correctly subtracted for non-SI maps.


2.6 COSMIC VARIANCE 18<br />

Non-zero ˜κ l estimated from a non-SI map will have contribution from the non-SI terms<br />

in full bias given in eq. (2.38). It is not inconceivable that for strong SI violation, B l<br />

over-corrects for the bias leading to negative values of ˜κ l . What is important is whether<br />

the measured ˜κ l differs from zero at a statistically significant level.<br />

2.6 Cosmic Variance<br />

Cosmic variance is defined as the variance of the estimator of the BiPS<br />

σ 2 = < ˜κ 2 l > − < ˜κ l > 2 (2.41)<br />

We can analytically compute the variance of ˜κ l using the Gaussianity of ∆T . Looking<br />

back at the eq.(2.30) we can see, we will have to calculate the eighth moment of the<br />

field<br />

〈∆T (ˆn 1 )∆T (ˆn 2 )∆T (ˆn 3 )∆T (ˆn 4 )∆T (ˆn 5 )∆T (ˆn 6 )∆T (ˆn 7 )∆T (ˆn 8 )〉. (2.42)<br />

Assuming Gaussianity of the field we can rewrite the eight point correlation in terms<br />

of two point correlations. One can write a simple code to do that 4 . This will give us<br />

(8 − 1)!! = 7 × 5 × 3 = 105 terms. These 105 terms consist of terms like:<br />

〈∆T (ˆn 1 )∆T (ˆn 2 )〉〈∆T (ˆn 3 )∆T (ˆn 4 )〉〈∆T (ˆn 5 )∆T (ˆn 6 )〉〈∆T (ˆn 7 )∆T (ˆn 8 )〉, (2.43)<br />

and all other permutations of them. On the other hand 〈˜κ l 〉 has a 4 point correlation<br />

in it which can also be expanded versus two point correlation functions. If we form<br />

〈˜κ 2 l 〉 − 〈˜κ l〉 2 , only 96 terms will be left which are in the following form<br />

( 2l + 1 ∫<br />

) 2 dΩ<br />

8π 2<br />

1 · · · dΩ 4 dRdR ′ χ l (R)χ l (R ′ )C(ˆn 1 , R ′ˆn 4 )C(ˆn 2 , R ′ˆn 3 )C(Rˆn 1 , ˆn 4 )C(Rˆn 2 , ˆn 3 )<br />

(2.44)<br />

and all other permutations. In Appendix A we explain how we can handle these 96<br />

terms to compute the following analytical expression for cosmic variance (details of the<br />

above calculations are presented in Appendix B.),<br />

σ 2 (˜κ SI<br />

l) = ∑<br />

]<br />

4 Cl 4 W l<br />

[2 4 (2l + 1)2<br />

2l + 1<br />

+ (−1)l (2l + 1) + (1 + 2(−1) l )Fll<br />

l<br />

l:2l≥l<br />

4 F90 software implementing this is available from the authors upon request


2.7 COSMIC VARIANCE 19<br />

+ ∑ ∑l+l 1<br />

l 1<br />

l 2 =|l−l 1 |<br />

4 C 2 l 1<br />

C 2 l 2<br />

W 2<br />

l 1<br />

W 2<br />

l 2<br />

[<br />

(2l + 1) + F<br />

l<br />

⎡<br />

+ 8 ∑ (2l + 1) 2<br />

2l 1 + 1 C2 l 1<br />

W 2 ⎣<br />

l 1<br />

l 1<br />

+ 16 (−1) l ∑<br />

l 1 :2l 1 ≥l<br />

Numerical computation of σ 2 SI<br />

(2l + 1) 2<br />

2l 1 + 1<br />

∑l+l 1<br />

l 2 =|l−l 1 |<br />

∑l+l 1<br />

l 2 =|l−l 1 |<br />

⎤2<br />

C l2 W l2<br />

⎦<br />

l1 l 2<br />

]<br />

C 3 l 1<br />

C l2 W 3<br />

l 1<br />

W l2 . (2.45)<br />

is fast. But the challenge is to compute Clebsch-Gordan<br />

coefficients for large quantum numbers. We use drc3j subroutine of netlib 5 in order to<br />

compute the Clebsch-Gordan coefficients in our codes. Again we can check the accuracy<br />

of our analytical estimation of cosmic variance by comparing it against the standard<br />

deviation of BiPS of 1000 simulations of SI <strong>CMB</strong> sky. The result is shown in Fig. 2.4<br />

and shows a very good agreement between the two.<br />

5<br />

100<br />

50<br />

0<br />

0<br />

-50<br />

-5<br />

-100<br />

5 10 15 20<br />

5 10 15 20<br />

Figure 2.4 Bias corrected ‘measurement’ of BiPS, ˜κ l for SI <strong>CMB</strong> maps <strong>with</strong> a best fit<br />

LCDM power spectrum smoothed by (left:) a two beam function W S<br />

l (20, 30) and (right:)<br />

a Gaussian beam Wl<br />

G (40). The cosmic error, σ(κ l ), obtained using 1000 independent<br />

realizations of <strong>CMB</strong> (full) sky map matches the analytical results shown by dotted curve<br />

<strong>with</strong> triangles . This shows a much better fit to the theoretical cosmic variance compared<br />

to what was obtained for 100 realizations Hajian & Souradeep (2003b)<br />

5 http://www.netlib.org/slatec/src/


2.7 SOURCES OF BREAKDOWN OF STATISTICAL ISOTROPY 20<br />

2.7 Sources of Breakdown of Statistical Isotropy<br />

An observed map of <strong>CMB</strong> anisotropy, ∆T obs (ˆn), is an n-dimensional vector (n =<br />

12 × 512 2 for the WMAP data at HEALPix resolution N side = 512). This observed<br />

map contains the true <strong>CMB</strong> temperature fluctuations, ∆T (ˆn), convolved <strong>with</strong> the beam<br />

and buried into noise and foreground contaminations. The observed map is related to<br />

the true map through this relation<br />

∆T obs = B∆T + n, (2.46)<br />

in which B is a matrix that contains the information about the beam smoothing effect<br />

and n is the contribution from instrumental noise and foreground contamination. Hence,<br />

the observed map is a realization of a Gaussian process <strong>with</strong> covariance C = C T + C N +<br />

C res where C T is the theoretical covariance of the <strong>CMB</strong> temperature fluctuations, C N<br />

is the noise covariance matrix and C res is the covariance of residuals of foregrounds.<br />

Breakdown of statistical isotropy can occur in any of these parts. In general we can<br />

divide these effects into two kinds:<br />

• Theoretical effects: These effects are theoretically motivated and are intrinsic to<br />

the true <strong>CMB</strong> sky, ∆T . We discuss three examples of these effects, i.e. non-trivial<br />

cosmic topology (Chapter 4), hidden patterns in the <strong>CMB</strong> maps which can be<br />

explained by models of homogeneous spaces known as Bianchi models (Chapter 5)<br />

and primordial magnetic fields, in the next section.<br />

• Observational artifacts: In an ideally cleaned <strong>CMB</strong> map, the true <strong>CMB</strong> temperature<br />

fluctuations are completely extracted from the observed map. But this<br />

is not always true. Sometimes there are some artifacts (related to B or n) left in<br />

the cleaned map which may in principle violate the SI. These effects are explained<br />

in Chapter 6.


l’<br />

l’<br />

2.8 PRIMORDIAL MAGNETIC FIELDS 21<br />

2.8 Primordial Magnetic Fields<br />

As the first example of theoretical motivations for models of breakdown of SI, we<br />

study the simplest configuration of cosmological magnetic fields. It has been shown<br />

that a cosmological magnetic field, generated during an early epoch of inflation [Ratra<br />

(1992); Bamba et al. (2004)], can generate <strong>CMB</strong> anisotropies [Durrer et al. (1998)].<br />

The presence of a preferred direction due to a homogeneous magnetic field background<br />

leads to non-zero off-diagonal elements in the covariance matrix [Chen et al. (2004)].<br />

This induces correlations between a l+1,m and a l−1,m multipole coefficients of the <strong>CMB</strong><br />

temperature anisotropy field in the following manner (see Fig. 2.5)<br />

20<br />

x 10 −3<br />

20<br />

x 10 −3<br />

18<br />

2.5<br />

18<br />

2.5<br />

16<br />

16<br />

2<br />

2<br />

14<br />

14<br />

12<br />

1.5<br />

12<br />

1.5<br />

10<br />

10<br />

8<br />

1<br />

8<br />

1<br />

6<br />

6<br />

4<br />

0.5<br />

4<br />

0.5<br />

2<br />

2<br />

2 4 6 8 10 12 14 16 18 20<br />

l<br />

0<br />

2 4 6 8 10 12 14 16 18 20<br />

l<br />

0<br />

Figure 2.5 Effect of a homogeneuos magnetic field on covariance matrix is the introduction<br />

of off-diagonal elements (right). In the absence of magnetic fields the covariance<br />

matrix is diagonal (left).<br />

〈a lm a ∗ l ′ m ′〉 = δ m,m ′[δ l,l ′C l + (δ l+1,l ′ −1 + δ l−1,l ′ +1D l )], (2.47)<br />

where D l is the power spectrum of off-diagonal elements of the covariance matrix. For<br />

a Harrison-Peebles-Yu-Zel’dovich scale-invariant spectrum, D l behaves as l −2 . More<br />

precisely, it is given by<br />

D l = 4 × 10 −16 l −2 ( B<br />

1nG )4 . (2.48)<br />

This clearly violates the statistical isotropy. If we substitute the above 〈a lm a ∗ l ′ m ′〉 into<br />

eqn. (2.22), we will get<br />

A lM<br />

l 1 l 2<br />

= δ M,0 [(−1) l 1√<br />

(2l1 + 1)C l1 δ l1 ,l 2<br />

δ l,0 (2.49)


2.8 PRIMORDIAL MAGNETIC FIELDS 22<br />

+ D l1<br />

∑<br />

m1=−l1,l1<br />

+ D l1<br />

∑<br />

m1=−l1,l1<br />

(−1) m 1<br />

C lM<br />

l 1 m 1 l 2 −m 1<br />

δ l1+1 ,l 2−1<br />

(−1) m 1<br />

Cl lM<br />

1 m 1 l 2 −m 1<br />

δ l1−1 ,l 2+1<br />

].<br />

The first line is the statistically isotropic part and it just contributes to κ 0 . But the<br />

interesting parts are the two other lines. We can substitute this expression into eqn.<br />

(2.25) and compute the BiPS predictions for magnetic fields (see fig. 2.6).<br />

7<br />

6<br />

B=16nG<br />

B=14nG<br />

B=12nG<br />

5<br />

4<br />

κ l<br />

3<br />

2<br />

1<br />

0<br />

1 2 3 4 5 6 7<br />

l<br />

Figure 2.6 BiPS predictions for primordial homogeneous magnetic fields.


Chapter 3<br />

SI Violation: Analytic Models<br />

In this chapter we model the SI violation using toy models of non-SI correlations<br />

which are analytically solvable. These toy models are motivated by various mechanisms<br />

of SI violation such as multiply connected spaces. Besides helping us to come to a good<br />

understanding of violation of SI, they shed a light on how BiPS method works. We will<br />

repeatedly refer to this chapter in the rest of this thesis to explain several signatures of<br />

different mechanisms of SI violation on BiPS.<br />

3.1 One Preferred Axis<br />

The simplest way of having an aniostropic correlation function is to introduce a<br />

preferred axis in the correlation space. The toy model we first consider has a preferred<br />

axis in the z direction and its two point correlation is given by<br />

C(ˆq 1 , ˆq 2 ) = 1 − ɛ (ˆq 1 · ẑ − ˆq 2 · ẑ) 2 , (3.1)<br />

where ɛ < 1/4 to ensure positivity of C(ˆq 1 , ˆq 2 ). This is an interesting case where the<br />

BiPS can be computed analytically. To get the analytical expression for BiPS in this<br />

toy model, we use eqn.(2.30) and evaluate the following integral<br />

∫ ∫ ∫<br />

κ l = (2l + 1) 2 1<br />

dΩ q1 dΩ q2 [ dR χ l (R) C(Rˆq<br />

8π 2 1 , Rˆq 2 )] 2 (3.2)<br />

∫ ∫ ∫<br />

= (2l + 1) 2 1<br />

dΩ q1 dΩ q2 [ dR χ l (R) {1 − ɛ (ˆq ′<br />

8π 2 1 · ẑ − ˆq′ 2 · ẑ)2 }] 2 .<br />

23


3.1 ONE PREFERRED AXIS 24<br />

In the above expression ˆq ′ i = R ωˆn ˆq i are the coordinates of the pixels after rotating them<br />

through an angle ω about an axis ˆn(Θ, Φ). Correlation only contains dot products of<br />

the pixel coordinates and the z axis, so one can equivalently calculate the function by<br />

keeping the directions ˆq 1 and ˆq 2 fixed and rotating the axes. In particular the ẑ axis<br />

rotated by an angle −ω about the same axis ˆn(Θ, Φ) goes to<br />

ẑ ′ = ẑ cos ω + ẑ(ˆn · ẑ)(1 − cos ω) − (ˆn × ẑ) sin ω (3.3)<br />

= (sin Φ cos Θ sin Θ(1 − cos ω) − sin Φ sin Θ sin ω,<br />

cos Θ sin Φ sin Θ(1 − cos ω) + cos Φ sin Θ sin ω,<br />

cos 2 Θ(1 − cos ω) + cos ω) .<br />

After substituting this in the correlation function we will get<br />

1 − ɛ (ˆq 1 · ẑ ′ − ˆq 2 · ẑ ′ ) 2 = 1 − ɛ((q 1x − q 2x ) 2 Z ′ 2<br />

x + (q 1y − q 2y ) 2 Z ′ 2<br />

y (3.4)<br />

+ (q 1z − q 2z ) 2 Z ′ 2<br />

z )<br />

−<br />

−<br />

−<br />

2ɛ (q 1x − q 2x )(q 1y − q 2y )Z ′ x Z ′ y<br />

2ɛ (q 1x − q 2x )(q 1z − q 2z )Z ′ xZ ′ z<br />

2ɛ (q 1y − q 2y )(q 1z − q 2z )Z ′ yZ ′ z<br />

Where Z i are components of vector ẑ, i. e. ẑ = (Z x , Z y , Z z ). We also introduce following<br />

shorthands for simplicity<br />

δ x = q 1x − q 2x (3.5)<br />

δ y = q 1y − q 2y<br />

δ z = q 1z − q 2z .<br />

The correlation function then becomes<br />

1 − ɛ (ˆq 1 · ẑ ′ − ˆq 2 · ẑ ′ ) 2 = 1 − ɛ(δ 2 xZ ′ 2<br />

x + δ 2 yZ ′ 2<br />

y + δ 2 zZ ′ 2<br />

z ) (3.6)<br />

−<br />

−<br />

−<br />

2ɛ δ x δ y Z ′ x Z ′ y<br />

2ɛ δ x δ z Z ′ x Z ′ z<br />

2ɛ δ y δ z Z ′ y Z ′ z


3.1 ONE PREFERRED AXIS 25<br />

Now we should substitute for z ′<br />

The ω dependence is only in z ′<br />

from eqn.(3.3) and evaluate the integrals of eqn(3.2).<br />

so it is better to do the ω integration first. We find that<br />

the cross terms like Z ′ x Z ′ z will vanish and of the non zero terms, the first term, 1, would<br />

only contribute to κ 0 because<br />

∫<br />

∫ π<br />

dR 1 χ l (R) = 4π dωχ l (ω) sin 2 ω/2 (3.7)<br />

0<br />

= 2π 2 δ l0<br />

which is a direct result of orthonormality of the characteristic function χ l (ω). The above<br />

result is in fact expected because the first term is the SI part of the correlation function<br />

and it sould only contribute to κ 0 . The anisotropic part would have a component of κ 2<br />

as well<br />

∫<br />

dR χ l (R) (ˆq 1 · ẑ ′ − ˆq 2 · ẑ ′ ) 2 = ξ(ˆq 1 , ˆq 2 )δ l0 + ζ(ˆq 1 , ˆq 2 )δ l2 (3.8)<br />

Putting everything together and computing the inner-most integral, we would ge<br />

∫<br />

dR χ l (R) C(Rˆq 1 , Rˆq 2 ) = 2π 2 δ l0 − ɛ(ξ(ˆq 1 , ˆq 2 )δ l0 + ζ(ˆq 1 , ˆq 2 )δ l2 ) (3.9)<br />

W here : ξ(ˆq 1 , ˆq 2 ) = 2π2<br />

3 (δ2 x + δ2 y + δ2 z )<br />

ζ(ˆq 1 , ˆq 2 ) = 2π2<br />

15 (2δ2 x − δ2 y − δ2 z ).<br />

And using these, we can compute the analytical expression for BiPS by carrying out<br />

this integral<br />

κ l =<br />

∫<br />

(2l + 1)2<br />

(8π 2 ) 2<br />

dΩ q1<br />

∫<br />

dΩ q2 [2π 2 δ l0 − ɛ(ξ(ˆq 1 , ˆq 2 )δ l0 + ζ(ˆq 1 , ˆq 2 )δ l2 )] 2 . (3.10)<br />

which will give<br />

where<br />

κ l =<br />

(2l + 1)2<br />

(8π 2 ) 2 [κ 0 δ l0 + κ 2 δ l2 ] , (3.11)<br />

κ 0 = 64<br />

27 π6 (27 − 36ɛ + 16ɛ 2 ) (3.12)<br />

κ 2 = 4096ɛ2 π 6<br />

.<br />

3375<br />

This exercise shows us the sensitivity of BiPS to the characteristic patterns of correlation.<br />

Any SI violating mechanism which causes a preferred direction in the correlation space,


3.2 THREE PREFERRED AXES 26<br />

will have a non-zero κ 2 .<br />

Examples of this are anisotropic toroidal topologies of the<br />

Universe and noise covariance matrices like WMAP noise which are studied in the next<br />

chapters and are shown to have a non-zero κ 2 .<br />

3.2 Three Preferred Axes<br />

Having seen the effect of one preferred axis on BiPS, we can consider a more general<br />

case in which we have three preferred directions in the correlation function. Each axis<br />

will cause a non-zero l = 2 component in BiPS. We attribute three strength factors ɛ 1 ,<br />

ɛ 2 and ɛ 3 to these directions to have a control on importance of each of them<br />

C(ˆq 1 , ˆq 2 ) = 1 − (ɛ 1 ((ˆq 1 − ˆq 2 ) · ˆx) 2 + ɛ 2 ((ˆq 1 − ˆq 2 ) · ŷ) 2 + ɛ 3 ((ˆq 1 − ˆq 2 ) · ẑ) 2 ) (3.13)<br />

to ensure positivity, we restrict our interest to models <strong>with</strong> ɛ i < 1/4. We can repeat the<br />

steps of our calculations in the last section to get the analytical expression for BiPS in<br />

this model. To avoid the repetition, we will only present the results of two important<br />

steps which will be useful later. The first is the integral over rotations which determines<br />

the non-zero l components<br />

∫<br />

dR χ l (R) C(Rˆq 1 , Rˆq 2 ) = 2π 2 δ l0 − (ξ(ˆq 1 , ˆq 2 )δ l0 + ζ(ˆq 1 , ˆq 2 )δ l2 ) (3.14)<br />

where<br />

ξ(ˆq 1 , ˆq 2 ) = 2π2<br />

3 (ɛ 1 + ɛ 2 + ɛ 3 )(δ 2 x + δ2 y + δ2 z ) (3.15)<br />

ζ(ˆq 1 , ˆq 2 ) = − 2π2<br />

15 ((−2ɛ 1 + ɛ 2 + ɛ 3 )δ 2 x + (ɛ 1 − 2ɛ 2 + ɛ 3 )δ 2 y + (ɛ 1 + ɛ 2 − 2ɛ 3 )δ 2 z ).<br />

Then we integrate over Ωˆq1 and Ωˆq2 and we get the BiPS for this toy model<br />

κ l =<br />

and κ 0 and κ 2 are given by the strength factors<br />

(2l + 1)2<br />

(8π 2 ) 2 [κ 0 δ l0 + κ 2 δ l2 ] , (3.16)<br />

κ 0 = 64<br />

27 π6 (27 + 16(ɛ 2 1 + ɛ2 2 + ɛ2 3 ) + 32(ɛ 1ɛ 2 + ɛ 1 ɛ 3 + ɛ 2 ɛ 3 ) − 36(ɛ 1 + ɛ 2 + ɛ 3 ))(3.17)<br />

= 64<br />

27 π6 (27 + 16Σ ɛ 2 + 32Π ɛ − 36Σ ɛ )<br />

κ 2 = 4096π6<br />

3375 (ɛ2 1 + ɛ 2 2 + ɛ 2 3 − ɛ 1 ɛ 2 − ɛ 1 ɛ 3 − ɛ 2 ɛ 3 )<br />

= 4096π6<br />

3375 (Σ ɛ 2 − Π ɛ)


3.3 OBLIQUE PREFERRED AXIS 27<br />

In the above expressions, we have used these conventions for simplicity<br />

Σ ɛ 2 = ɛ 2 1 + ɛ 2 2 + ɛ 2 3 (3.18)<br />

Σ ɛ = ɛ 1 + ɛ 2 + ɛ 3<br />

Π ɛ = ɛ 1 ɛ 2 + ɛ 1 ɛ 3 + ɛ 2 ɛ 3<br />

Interestingly we can see that the only non-zero component is l = 2. If the strength<br />

factors are the same for three three preferred axes, κ 2 will vanish and we will get a null<br />

BiPS. This is expectable because this is in fact a statistically isotropic model which can<br />

be seen explicitly from the two point correlation<br />

C(ˆq 1 , ˆq 2 ) = 1 − ɛ [ ((ˆq 1 − ˆq 2 ) · ˆx) 2 + ((ˆq 1 − ˆq 2 ) · ŷ) 2 + ((ˆq 1 − ˆq 2 ) · ẑ) 2] (3.19)<br />

= 1 − ɛ(ˆq 1 − ˆq 2 ) · (ˆq 1 − ˆq 2 )<br />

= (1 + 2ɛ) − 2ɛ(ˆq 1 · ˆq 2 )<br />

= C(ˆq 1 · ˆq 2 ).<br />

3.3 Oblique Preferred Axis<br />

Three axes in the previous toy model where perpendicular to each other. A more<br />

general case is where the axes make arbitrary angles to each other. This model can<br />

be obtained from the previous one by the act of a linear transform that changes the<br />

direction of the preferred axes;<br />

⎛ ⎞<br />

x<br />

⎜<br />

⎝ y ⎟<br />

⎠ −→R<br />

z<br />

⎛ ⎞<br />

x R<br />

⎜<br />

⎝ y R ⎟<br />

⎠ (3.20)<br />

z R<br />

A convenient parametrization of this transform matrix is<br />

R =<br />

⎛<br />

⎞<br />

1 cos α 1 cos α 2<br />

⎜<br />

⎝ 0 sin α 1 cos α 3 sin α ⎟ 2 ⎠ (3.21)<br />

0 0 sin α 2 sin α 3


3.3 OBLIQUE PREFERRED AXIS 28<br />

After this transformation, three preferred directions will make angles α 1 , α 2 and α 3 <strong>with</strong><br />

each other. The toy correlation fucnation <strong>with</strong> these preferred directions may be given<br />

as<br />

C(ˆq 1 , ˆq 2 ) = 1 − [ ɛ 1 ((ˆq 1 − ˆq 2 ) · ˆx R ) 2 + ɛ 2 ((ˆq 1 − ˆq 2 ) · ŷ R ) 2 + ɛ 3 ((ˆq 1 − ˆq 2 ) · ẑ R ) 2] (3.22)<br />

Integral over the rotation group elements would tell us l = 2 is the (nontrivial) non-zer<br />

component<br />

∫<br />

dR χ l (R) C(Rˆq 1 , Rˆq 2 ) = 2π 2 δ l0 − (ξ(ˆq 1 , ˆq 2 )δ l0 + ζ(ˆq 1 , ˆq 2 )δ l2 ) (3.23)<br />

in which<br />

ξ(ˆq 1 , ˆq 2 ) = 2π2<br />

3 Σ ɛ(δ 2 x + δ2 y + δ2 z ) (3.24)<br />

ζ(ˆq 1 , ˆq 2 ) = π2<br />

15 [Σ ɛ + 3(ɛ 1 + ɛ 2 cos 2α 1 + ɛ 3 cos 2α 2 )] δ 2 x<br />

−<br />

−<br />

π2<br />

60 [8ɛ 1 − 4ɛ 2 (1 − 3 cos 2α 1 ) + 6ɛ 3 (cos 2α 2 − cos 2α 3 + cos 2α 2 cos 2α 3 )] δ 2 y<br />

π2<br />

60 [8Σ ɛ + 6ɛ 3 (−1 + cos 2α 2 + cos 2α 3 − cos 2α 2 cos 2α 3 )] δ 2 z<br />

+ 4π2<br />

5 [ɛ 2 cos α 1 sin α 1 + ɛ 3 cos α 2 cos α 3 sin α 2 ] δ x δ y<br />

+ 4π2 [ ]<br />

ɛ3 cos α 3 sin 2 α 2 sin α 3 δy δ z + 2π2<br />

5<br />

5 [ɛ 3 sin 2α 2 sin α 3 ] δ x δ z<br />

This is the most general case. For simplicity we can confine ourselves to models <strong>with</strong><br />

only one oblique axis and two orthogonal ones,<br />

α 2 = α 3 = π/2. (3.25)<br />

BiPS for this model can then be computed by integrating the square of the above<br />

expression over all possible directions on the sky<br />

κ l =<br />

(2l + 1)2<br />

(8π 2 ) 2 [κ 0 δ l0 + κ 2 δ l2 ] , (3.26)<br />

which shows that a non-zero κ 2 is the signature of these models, where<br />

κ 0 = 64<br />

27 π6 (27 + 16Σ ɛ 2 + 32Π ɛ − 36Σ ɛ ) (3.27)<br />

κ 2 = 2048π6<br />

3375 (2Σ ɛ 2 − 2Π ɛ + 3ɛ 1 ɛ 2 (cos α 1 + 1))


3.5 TOY MODEL WITH CROSS TERMS 29<br />

κ 2 varies as cos α 1 . Interestingly, if we set all three strength factors to be equal, say<br />

ɛ 3 = ɛ 2 = ɛ 1 (3.28)<br />

we will still have a non-zero l = 2 components in BiPS which is given as<br />

κ 0 = 64 3 π6 (3 − 12ɛ 1 + 16ɛ 2 1) (3.29)<br />

κ 2 = 4096π6<br />

1125 ɛ2 1 cos 2 α 1 .<br />

κ 2 given above is proportional to cos 2 α 1 and tends to zero when α 1 → π/2 and the<br />

correlation function tends to a SI correlation function.<br />

3.4 Toy Model With Cross Terms<br />

form<br />

Sometimes in our analytical calculations of BiPS, we encounter cross terms of thsi<br />

f(ˆq 1 , ˆq 2 ) = ((ˆq 1 − ˆq 2 ) · ˆx)((ˆq 1 − ˆq 2 ) · ŷ) (3.30)<br />

Which has two preferred axes in ˆx and ŷ directions and is of the first order in each of<br />

them. Although this is not an example of a physical two point function, it is useful to<br />

know the analytical BiPS for that (we will use this in our later calcularions). Integral<br />

over rotations of the above function will give us a non-zero l = 2 component<br />

∫<br />

dR χ l (R) f(Rˆq 1 , Rˆq 2 ) = 2π2<br />

5 δ xδ y δ l2 (3.31)<br />

This tells us that each cross term will cause a non-zero l = 2 component. Using this,<br />

we can conclude that if there are three cross terms <strong>with</strong> strength factors ɛ 1 , ɛ 2 and ɛ 3 ,<br />

the result of the above integral would be a superposition of three terms<br />

∫<br />

dR χ l (R) f(Rˆq 1 , Rˆq 2 ) = 2π2<br />

5 (ɛ 1δ x δ y + ɛ 2 δ x δ y + ɛ 3 δ y δ z )δ l2 (3.32)<br />

This has no l = 0 component because there is no SI part in our toy model. BiPS of this<br />

model can be calculated to be<br />

κ l =<br />

(2l + 1)2 1024π 6 Σ ɛ 2<br />

δ<br />

(8π 2 ) 2 l2 , (3.33)<br />

1125<br />

This is always non-zero even for equal strength factors, and tends to zero as Σ ɛ 2 → 0.


3.5 FOURTH ORDER TOY MODELS WITH THREE AXES 30<br />

3.5 Fourth Order Toy Models With Three Axes<br />

We showed that all anisotropic second order correlation functions lead to a non-zero<br />

κ 2 in BiPS. Next toy model we consider is a fourth order toy model <strong>with</strong> three preferred<br />

directions along the ˆx, ŷ and ẑ directions <strong>with</strong> strength factors ɛ 1 , ɛ 2 and ɛ 3 respectively,<br />

C(ˆq 1 , ˆq 2 ) = 1 − (ɛ 1 ((ˆq 1 − ˆq 2 ) · ˆx) 4 + ɛ 2 ((ˆq 1 − ˆq 2 ) · ŷ) 4 + ɛ 3 ((ˆq 1 − ˆq 2 ) · ẑ) 4 ) (3.34)<br />

Integration over rotations of this correlation function will cause a non-zero l = 2 component<br />

as well as a non-zero l = 4 one,<br />

∫<br />

dR χ l (R) C(Rˆq 1 , Rˆq 2 ) = 2π 2 δ l0 − (ξ(ˆq 1 , ˆq 2 )δ l0 + ζ(ˆq 1 , ˆq 2 )δ l2 + η(ˆq 1 , ˆq 2 )δ l4 ) (3.35)<br />

W here : ξ(ˆq 1 , ˆq 2 ) = 2π2<br />

5 Σ ɛ(δ 2 x + δ 2 y + δ 2 z) 2<br />

ζ(ˆq 1 , ˆq 2 ) = − 4π2<br />

35 ((−2ɛ 1 + ɛ 2 + ɛ 3 )δ 2 x + (ɛ 1 − 2ɛ 2 + ɛ 3 )δ 2 y + (ɛ 1 + ɛ 2 − 2ɛ 3 )δ 2 z),<br />

η(ˆq 1 , ˆq 2 ) = 2π2<br />

315 (δ4 x(8ɛ 1 + 3ɛ 2 + 3ɛ 3 ) + δ 4 y(3ɛ 1 + 8ɛ 2 + 3ɛ 3 )<br />

+ δ 4 z (3ɛ 1 + 3ɛ 2 + 8ɛ 3 ) − 6δ 2 x δ2 y (4ɛ 1 + 4ɛ 2 − ɛ 3 )<br />

− 6δ 2 y δ2 z (−ɛ 1 + 4ɛ 2 + 4ɛ 3 ) − 6δ 2 x δ2 z (4ɛ 1 − ɛ 2 + 4ɛ 3 )).<br />

Integrating the above expression is a tedious job. To simlify our calculations we can<br />

consider a special case in which three strength factors are the same<br />

ɛ 3 = ɛ 2 = ɛ 1 (3.36)<br />

This simplifies the calculations immensely and we can easily get the BiPS for this class<br />

of toy models to be<br />

κ l =<br />

(2l + 1)2<br />

(8π 2 ) 2 [κ 0 δ l0 + κ 2 δ l2 + κ 4 δ l4 ] , (3.37)<br />

where<br />

κ 0 = 64<br />

125 π6 (125 − 800ɛ 1 + 2304ɛ 2 1) (3.38)<br />

κ 2 = 0<br />

κ 4 = 262144<br />

212625 π6 ɛ 2 1


3.7 TOY MODEL II 31<br />

In this case, the l = 2 component of the power spectrum vanishes. This result along<br />

<strong>with</strong> what we obtained in 3.17 will hint us an interesting property of BiPS. Three<br />

perpendicular preferred directions in the correlation function lead to a zero κ 2 , but if<br />

they appear in higher orders in the correlation function, they will cause non-zero κ 4 and<br />

as we will see later in higher orders may cause higher l components non-zero.<br />

3.6 Toy Model II<br />

We can consider another toy correlation which is analytically solvable and has preferred<br />

directions in the following way<br />

C(ˆq 1 , ˆq 2 ) = (1 + ɛˆq 1 · ẑ)(1 + ɛˆq 2 · ẑ), (3.39)<br />

where ɛ < 1 ensures positivity of the correlation function.<br />

Integrating over rotations of the two point correlation will lead to the following result<br />

∫<br />

dR χ l (R) C(Rˆq 1 , Rˆq 2 ) = 2π 2 δ l0 − ɛ(ξ(ˆq 1 , ˆq 2 )δ l0 + ζ(ˆq 1 , ˆq 2 )δ l2 ) (3.40)<br />

W here : ξ(ˆq 1 , ˆq 2 ) = 2π2<br />

3 (δ2 x + δ2 y + δ2 z )<br />

ζ(ˆq 1 , ˆq 2 ) = 2π2<br />

15 (2δ2 x − δ2 y − δ2 z ).<br />

And BiPS is computed by integrating the square of the above expression over all possible<br />

pairs of pixels on the sky,<br />

κ l =<br />

(2l + 1)2<br />

(8π 2 ) 2 [κ 0 δ l0 + κ 1 δ l1 + κ 2 δ l2 ]. (3.41)<br />

Where , κ 0 κ 1 and κ 2 are given by<br />

κ 0 = 16<br />

27 4π6 (27 + ɛ 4 ), (3.42)<br />

κ 1 = 32<br />

27 π6 ɛ 2 ,<br />

κ 2 = 32<br />

675 π6 ɛ 4 .<br />

These show that BiPS should have three non-zero components which are κ 0 , κ 1 and κ 2 .<br />

The rest of the components are zero. This is the first (and only) analytical toy model<br />

which has a non-zero odd-l component of BiPS.


3.7 TOROIDAL SPACES 32<br />

3.7 Toroidal Spaces<br />

A family of interesting models in which the SI condition is violated is the family of<br />

compact spaces. As we will see in Chapter 4, this breakdown of statistical isotropy is<br />

equivalent to having strongly correlated spots which are physically very far from each<br />

other. This is a hint of the existance of preferred directions in the correlation spaces of<br />

these models. In Chapter 4 we compute C(ˆq, ˆq ′ ) for <strong>CMB</strong> anisotropy in toroidal spaces<br />

using regularized method of images Bond, Pogosyan & Souradeep (1998, 2000). The<br />

BiPS can be numerically computed for these models from their two point correlation<br />

matrices. The results can be understood (and predicted) using the leading order terms<br />

of the correlation function in a torus where κ l can be calculated analytically. These are<br />

discussed in the next chapter. And as we will see, they show that BiPS is a very useful<br />

tool in studying the topology of the universe.


Chapter 4<br />

Signatures of Cosmic Topology<br />

The question of size and the shape of our universe are very old problems that have<br />

received increasing attention over the past few years, Ellis (1971); Lachieze-Rey &<br />

Luminet (1995); Levin et al. (1998); de Oliveira-Costa et al. (1996); Bond, Pogosyan &<br />

Souradeep (1998, 2000); Gomero et al. (2002); de Oliveira-Costa et al. (2003); Dineen<br />

et al. (2004); Copi et al. (2004); Mota et al. (2004). Although a multiply connected<br />

universe sounds non-trivial, but there are theoretical motivations, Linde (2004); Levin<br />

(2002), to favor a spatially compact universe. One possibility to have a compact flat<br />

universe is the consideration of multiply connected spaces. The oldest way of searching<br />

for global structure of the universe is by identifying ghost images of local galaxies and<br />

clusters or quasars at higher redshifts, Gomero et al. (1998, 2000, 2001) and Lachieze-<br />

Rey & Luminet (1995). This method can probe the topology of the universe only<br />

on scales substantially smaller than the apparent radius of the observable universe.<br />

Another method to search for the shape of the universe is through the effect on the<br />

power spectrum of cosmic density perturbation fields. In principle, this effect can be<br />

observed in the distribution of matter in the universe and the <strong>CMB</strong> anisotropy. Over the<br />

past few years, many independent methods have been proposed to search for evidence of<br />

a finite universe in <strong>CMB</strong> maps. These methods can be classified in three main groups.<br />

• Using the angular power spectrum of <strong>CMB</strong> anisotropies to probe the topology of<br />

the Universe. The angular power spectrum, however, is inadequate to characterize<br />

the peculiar form of the anisotropy manifest in small universes of this type. It has<br />

33


4.1 STATISTICAL ISOTROPY VIOLATION IN COMPACT SPACES 34<br />

been shown that a nontrivial topology breaks the SI down and as a result, there<br />

is more information in a map of temperature fluctuations than just the angular<br />

power spectrum, Levin et al. (1998); Bond, Pogosyan & Souradeep (1998, 2000);<br />

Hajian & Souradeep (2003a).<br />

• The second class of methods are direct methods which rely on multiple imaging of<br />

the <strong>CMB</strong> sky. The most well known methods amongst these methods are S-map<br />

statistics, de Oliveira-Costa et al. (1996, 2003) and the search for circles-in-thesky,<br />

Cornish, Spergel & Starkman (1998); Cornish et al. (2003). In Appendix C<br />

we study the search for planar symmetries in <strong>CMB</strong> anisotropy maps.<br />

• Third class of methods are indirect probes which deal <strong>with</strong> the correlation patterns<br />

of the <strong>CMB</strong> anisotropy field by using an appropriate combination of coefficients<br />

of the harmonic expansion of the field, Dineen et al. (2004); Donoghue et al.<br />

(2004); Hajian & Souradeep (2003a); Copi et al. (2004); Ferreira & Magueijo<br />

(1997).<br />

In the next section we show that SI violation is the smoking gun of compact spaces and<br />

can be used to detect and constrain these spaces. The strength of this method lies in the<br />

fact that it works even if the fundamental domain is slightly bigger than the diameter<br />

of the surface of last scattering, in which case the identified circles do not exist.<br />

4.1 Statistical Isotropy Violation in Compact Spaces<br />

One of the most important properties of compact spaces is that statistical isotropy<br />

is broken in them. Here we follow Riazuelo et al. (2004a) to calculate the covariance<br />

matrix in the universal cover and the quotient spaces in general. We use this result to<br />

show how one can use this property to search for compact spaces <strong>with</strong> the use of BiPS.


4.1 STATISTICAL ISOTROPY VIOLATION IN COMPACT SPACES 35<br />

Two point correlations in the universal cover – a general case<br />

The universal cover is a space of constant curvature and hence can be E 3 , H 3 or S 3 .<br />

The metric of the universal cover is a FLRW metric<br />

ds 2 = −c 2 dt 2 + a 2 (t) [ dχ 2 + f 2 K(χ)dΩ 2] , (4.1)<br />

where dΩ 2 = dθ 2 + sin 2 θ dϕ 2 , and χ is the comoving radial distance. f K (χ) for different<br />

geometries is given by<br />

⎧<br />

⎪⎨<br />

sinh( √ |K| χ)/ √ |K| (hyperbolic case, K < 0)<br />

f K (χ) = χ (flat case, K = 0)<br />

⎪⎩<br />

sin( √ K χ)/ √ K (spherical case, K > 0)<br />

(4.2)<br />

We usually work in Fourier space because it is more convenient to work <strong>with</strong> the individual<br />

modes of our system which evolve independently. The allowed modes are the<br />

eigenmodes of the space, Y [X]<br />

k lm<br />

(where superscript [X] shows the space of interest) given<br />

by (for example) Laplace equation<br />

∇ 2 Y [X]<br />

k lm + (k2 − K)Y [X]<br />

k lm<br />

= 0. (4.3)<br />

Eigenmodes Y [X]<br />

k lm<br />

can be decomposed into a radial part and an angular part like this<br />

Y [X]<br />

k lm = R[X] kl<br />

(χ)Y lm (ˆn) (4.4)<br />

To compute the observables of <strong>CMB</strong> anisotropy, we start from a set of initial conditions<br />

which is given to us by an early universe scenario. Modes that we study, evolve before<br />

they enter the Hubble radius till now. From this we can compute the ∆T/T at each<br />

point on the sky which is a random variable, and hence we must study its statistics<br />

which is assumed to be Gaussian. To do this, here we will first consider a flat case and<br />

then we will generalize our result to all cases. In a flat case the computation of the<br />

<strong>CMB</strong> anisotropies will contain three major steps<br />

(i)<br />

eigenmodes of Laplacian are plain waves given by<br />

e i(k·x) ,<br />

hence the Fourier modes of temperature fluctuations are<br />

∫<br />

∆T<br />

T (ˆn) = d 3 k ∆T<br />

ei(k·x)<br />

(2π)<br />

3/2<br />

T (ˆk)


4.1 STATISTICAL ISOTROPY VIOLATION IN COMPACT SPACES 36<br />

(ii)<br />

initial conditions are given by the power spectrum P Φ (k) of primordial fluctuations<br />

in the gravitational potential field Φ. These are then mapped to<br />

<strong>CMB</strong> anisotropies in k-space,<br />

∆T<br />

T (ˆk) ∝ √ P Φ (k)ê k ,<br />

where in the simplest inflationary models, ê k is a three dimensional Gaussian<br />

random variable, which satisfies<br />

〈ê k ê ∗ k〉 = δ(k − k ′ )<br />

In simply connected spaces, reality condition of the temperature field dictates<br />

that<br />

ê −k = ê ∗ k.<br />

(iii)<br />

(iv)<br />

initial condition are mapped (related) to the present temperature fluctuations<br />

in a given direction on the sky, ˆn, by a linear convolution operator<br />

O [E3 ]<br />

k<br />

the result of all this can be summarized as<br />

∫<br />

∆T<br />

T (ˆn) = d 3 k ]<br />

(2π) 3/2 O[E3 k<br />

e i(k·x)√ P Φ (k)ê k . (4.5)<br />

Using the addition theorem<br />

and<br />

l∑<br />

m=−l<br />

Y lm (ˆn)Y ∗<br />

lm (ˆn′ ) = 2l + 1<br />

4π P l(cos α), (4.6)<br />

e ik·r =<br />

∞∑<br />

(2l + 1)i l j l (kr)P l (θ k,r ), (4.7)<br />

l=0<br />

eqn. (4.5) can be written as<br />

∫<br />

∆T<br />

T (ˆn) = k 2 dk √ P Φ (k) O [E3 ] 1 ∑<br />

k<br />

i l j l (kχ)Y lm (ˆn)Ylm ∗<br />

4π<br />

(ˆk)ê k (4.8)<br />

l,m<br />

= ∑ ∫<br />

i l k 2 dk √ (√ ) [∫ ]<br />

P Φ (k) O [E3 ] 2<br />

k<br />

π j l(kχ) dΩ k Ylm ∗ (ˆk)ê k Y lm (ˆn)<br />

l,m<br />

= ∑ ∫<br />

i l k 2 dk √ ( )<br />

P Φ (k) O [E3 ]<br />

k<br />

Y [E3 ]<br />

k lm<br />

ê lm (k), (4.9)<br />

l,m


4.1 STATISTICAL ISOTROPY VIOLATION IN COMPACT SPACES 37<br />

where ê lm (k) is the spherical harmonic transform of ê k and hence obeys the isotropy<br />

condition<br />

〈ê lm (k)ê ∗ l ′ m ′(k′ )〉 = 1 k 2 δ(k − k′ )δ ll ′δmm ′ . (4.10)<br />

Eqn. (4.9) can be written as<br />

∆T<br />

T (ˆn) = ∑ ∫<br />

i l<br />

l,m<br />

k 2 dk √ P Φ (k) O [E3 ]<br />

k<br />

R [E3 ]<br />

k l<br />

ê lm (k)Y lm (ˆn), (4.11)<br />

and a lm can be read from that to be<br />

a lm = i l ∫<br />

k 2 dk √ P Φ (k) G l (k)ê lm (k) (4.12)<br />

where<br />

G l (k) = O [E3 ]<br />

k<br />

R [E3 ]<br />

k l<br />

and is computed from the local physics of the fluctuations.<br />

This was all for flat spaces. But for non-flat cases, everything is exactly the same,<br />

except that the eigenmodes are different. Therefor while working <strong>with</strong> non-flat spaces,<br />

there are two points to be taken care of. First, j l must be replaced by ultra-spherical<br />

Bessel functions Zaldarriaga et al. (1998). And second, in closed spaces, the modes are<br />

discrete, so we should replace the integral by a discrete sum over eigenmodes.<br />

However, the observables of <strong>CMB</strong> anisotropy are not a lm but their variance. In the<br />

case of above spaces (E 3 , H 3 and S 3 ), which are statistically isotropic, the covariance<br />

matrix turns out to be diagonal <strong>with</strong> angular power spectrum on its diagonals<br />

〈a lm a ∗ l ′ m ′〉 = C lδ ll ′δ mm ′ (4.13)<br />

which means that the angular power spectrum C l is the trace of the covariance matrix,<br />

C ll′<br />

mm ′ = 〈a lma ∗ l ′ m ′〉, C l = 1<br />

2l + 1<br />

that is<br />

C l =<br />

∫<br />

∑<br />

Cmm ll (4.14)<br />

m<br />

k 2 dk<br />

(<br />

G [X]<br />

l<br />

(k)) 2<br />

PΦ (k). (4.15)


4.1 STATISTICAL ISOTROPY VIOLATION IN COMPACT SPACES 38<br />

Effect of topology<br />

Imposing the topology, will not change the local physics of our problem. Because all<br />

physical equations (including cosmological equations which are derived from Einstein<br />

equations) are local differential equations and are untouched by the global properties of<br />

the space. This implies that implementing the topology will not alter O [X]<br />

k<br />

in the above<br />

calculation. But it will change the modes that can exist in the universe. One should<br />

solve the Laplace equation in the compact space X/Γ in order to find the eigenmodes<br />

of Laplacian, Ψ [Γ]<br />

ks ,<br />

∇ 2 Ψ [Γ]<br />

ks + (k2 − K)Ψ [Γ]<br />

ks<br />

= 0. (4.16)<br />

These eigenmodes are given by the eigenmodes of the universal covering space as<br />

Ψ [Γ]<br />

ks = ∞<br />

∑<br />

l∑<br />

l=0 m=−l<br />

ξ [Γ]s<br />

klm Y[X] k lm<br />

(4.17)<br />

where s describes the degenerate eigenmodes and k is discrete. One should then use<br />

these eigenmodes and use a discrete sum instead of integral in eqn. (4.9) in order to<br />

calculate the temperature anisotropies in compact spaces. In a quotient space of volume<br />

V this becomes<br />

∆T<br />

T<br />

(ˆn) =<br />

(2π)3<br />

V<br />

∑<br />

k,s<br />

( ) √<br />

O [X]<br />

k<br />

Ψ [Γ]<br />

ks PΦ (k)ê k (4.18)<br />

we can denote ê k by ê ks and the orthogonality conditition will become<br />

〈ê k ê ∗ k 〉 = V<br />

(2π) 3 δ kk ′δ ss ′ (4.19)<br />

Substituting eqn. (4.17) in eqn. (4.18), we will find the analytical expression for temperature<br />

fluctuations in a compact space<br />

∆T<br />

T<br />

(ˆn) =<br />

(2π)3<br />

V<br />

from which a lm can be read off to be<br />

a lm = (2π)3<br />

V<br />

∑ ∑<br />

∑<br />

k<br />

k,s<br />

l,m<br />

( √<br />

ξ [Γ]s<br />

klm O[X] k<br />

Y [X]<br />

k lm)<br />

PΦ (k)ê ks (4.20)<br />

√<br />

PΦ (k)O [X]<br />

k<br />

(<br />

R [X]<br />

kl<br />

) ∑<br />

s<br />

ξ [Γ]s<br />

klmêks (4.21)<br />

The covariance matrix is calculated using the above expression, and is given by<br />

〈a lm a ∗ (2π)3 ∑ ( ) ( ) ∑<br />

l ′ m ′〉 = P Φ (k)O [X]<br />

k<br />

R [X]<br />

kl<br />

O [X]<br />

k<br />

R [X]<br />

kl<br />

ξ [Γ]s<br />

V<br />

′ klm ξ[Γ]s kl ′ m<br />

(4.22)<br />

′<br />

k<br />

s


4.2 METHOD OF IMAGES 39<br />

We should notice the existence of the term ∑ s ξ[Γ]s klm ξ[Γ]s kl ′ m ′ which is not necessarily diagonal<br />

in l and m.<br />

This is the statement of breakdown of statistical isotropy in compact<br />

spaces. As we mentioned earlier, the angular power spectrum which is given by C l =<br />

1<br />

2l+1<br />

∑<br />

m Cll mm<br />

does not have the whole information of the temperature field. Also as<br />

we showed in A, breakdown of SI imposes large cosmic variance on the angular power<br />

spectrum. All together, these make the angular power spectrum not a convenient probe<br />

of the cosmic topology. We use the BiPS which exploits all the information in the field<br />

by combining the off-diagonal elements in the covariance matrix into a power spectrum<br />

A lM<br />

ll ′<br />

= ∑ mm ′ (−1) m′ C lM<br />

lml ′ −m ′Cll′ mm ′,<br />

κ l = ∑<br />

l,l ′ ,M<br />

|A lM<br />

ll ′ |2 . (4.23)<br />

However this is not the fastest way of computing two point correlations in compact<br />

spaces.<br />

In the next chapter we study another method which is very convenient for<br />

computing two point correlations and generating simulated <strong>CMB</strong> skies in any compact<br />

space.<br />

4.2 Method of images<br />

Another method of computing two point correlations in compact spaces is the regularized<br />

method of images which was introduced and developed by Bond, Pogosyan &<br />

Souradeep (1998, 2000). The idea of the method of images is that since compact spaces<br />

are obtained from tilings of the fundamental domain, one should take all the images of<br />

each point into account while computing the two point correlations. This method saves<br />

us from computing the eigenmodes in compact spaces which is usually a numerically<br />

tedious job. Method of images has a nice mathematical origin which is given below.<br />

Like previous chapter we expand the temperature anisotropies in any space in terms<br />

of a complete, orthonormal set of functions in that space, e.g. eigenfunctions of the<br />

Laplacian operator denoted by Ψ [Γ]<br />

ks<br />

(∇ 2 + k 2 − K)Ψ [Γ]<br />

ks = 0.<br />

The spectrum of the Laplacian on a compact manifold is discrete. Hence the two point


4.2 METHOD OF IMAGES 40<br />

correlations will be given by<br />

ξ(x, x ′ ) = ∑ k,s<br />

Ψ [Γ]<br />

ks (x)Ψ[Γ] ks (x′ )P Φ (k). (4.24)<br />

Except in simple spaces, neither the spectrum of Laplacian {k}, nor the eigenfunctions<br />

Ψ [Γ]<br />

ks<br />

(x) are known and thus eqn. 4.24 can not be used to compute the two point correlation<br />

on compact spaces in general. In contrast the universal cover, M [X] , of the<br />

compact space is usually a very simple manifold whose eigenmodes are very well studied<br />

and two point correlations on these spaces (which are E 3 , S 3 or H 3 ) are usually very easily<br />

computable. The regularized method of images allows computation of the correlation<br />

function ξ [Γ] (x, x ′ ) on a compact space 1 M [Γ] = M [X] /Γ. The eigenfunctions Ψ [Γ]<br />

ks (x)<br />

on M [Γ] are eigenfunctions on the universal cover, M [X] too. So, they simultaneously<br />

satisfy the following integral equations for eigenfunctions on manifolds M [Γ] and M [X]<br />

∫<br />

dx ′ ξ [X] (x, x ′ )Ψ [Γ]<br />

ks (x′ ) = P Φ (k)Ψ [Γ]<br />

ks<br />

(x), (4.25)<br />

∫M [X] dx ′ ξ [Γ] (x, x ′ )Ψ [Γ]<br />

ks (x′ ) = P Φ (k)Ψ [Γ]<br />

ks (x).<br />

M [Γ]<br />

The eigenfunctions Ψ [Γ]<br />

ks<br />

(x) are automorphic <strong>with</strong> respect to Γ<br />

Ψ [Γ]<br />

ks<br />

(γx) = Ψ[Γ] (x) ∀γ ∈ Γ. (4.26)<br />

ks<br />

Using eqns (4.25) and (4.26) and the fact that M [Γ] tessellates M [X] , we get<br />

∫<br />

∫<br />

dx ′ ξ [Γ] (x, x ′ )Ψ [Γ]<br />

ks (x′ ) = dx ′ ξ [X] (x, x ′ )Ψ [Γ]<br />

ks (x′ ) (4.27)<br />

M [Γ] M [X]<br />

= ∑ ∫<br />

dx ′ ξ [Γ] (x, γx ′ )Ψ [Γ]<br />

ks (x′ )<br />

γ∈Γ<br />

M [Γ]<br />

∫ [ ]<br />

∑<br />

= dx ′ ξ [Γ] (x, γx ′ ) Ψ [Γ]<br />

ks (x′ ).<br />

M [Γ] γ∈Γ<br />

Since Ψ [Γ]<br />

ks<br />

(x) are orthonormal functions and the above equation should be satisfied for<br />

every x, we conclude that<br />

ξ [Γ] (x, x ′ ) = ∑ γ i ∈Γ<br />

ξ [X] (x, γ i x ′ ), (4.28)<br />

1 The compact manifold M [Γ] can be regarded as a quotient of a simply connected manifold M [X]<br />

by a fixed point free discontinuous subgroup Γ of the isometry group of M [X] . The elements γ ∈ Γ<br />

generate images of M [Γ] which tessellate the universal cover M u .


4.3 A BIPOLAR POWER SPECTRUM STUDY OF COSMIC TOPOLOGY 41<br />

This is the main equation of the method of images, which implies that the correlation<br />

function on a compact space is given by the sum over the correlation function on its<br />

universal cover calculated between x and the images γ i x ′ . But the sum in this equation<br />

can be obtained analytically in a closed form only in a few cases, e.g. the simple flat<br />

torus (which we have already seen it). For a numerical implementation of the sum over<br />

images, it is useful to present the formal expression (4.28) as the limit of a sequence of<br />

partial sums<br />

ξ [Γ] (x, x ′ ) = lim<br />

N→∞<br />

N∑ [<br />

ξ [X] (x, γ i x ′ ) ] , (4.29)<br />

i=0<br />

<strong>with</strong> the discrete motions γ i sorted in increasing separation, d(x, γ i x ′ ) ≤ d(x, γ i+1 x ′ ).<br />

In practice, the summation over images is carried out over a sufficiently large but finite<br />

number of images, from which the limit N → ∞ is estimated. Accuracy is enhanced<br />

if the γ i used correspond to a tessellation of M [X] <strong>with</strong> the base-point of the Dirichlet<br />

domain shifted to x. In this case the Nth partial term of eq.(4.28) is equal to<br />

ξ [Γ] (x, x ′ ) |N =<br />

N∑<br />

ξ [X] (x, γ i x ′ ). (4.30)<br />

i=0<br />

A more effective and simpler limiting procedure for implementing the regularized<br />

method of images is to explicitly sum images up to a radius r ∗ .<br />

This further eliminates<br />

the need to know the precise shape of Dirichlet domains (and integrate ξ [X] over<br />

potentially complicated shapes).<br />

⎡<br />

⎤<br />

ξ [Γ] = lim ⎣ ∑<br />

ξ [X] (r j ) ⎦ , r j = d(x, γ j x ′ ) ≤ r j+1 . (4.31)<br />

r ∗→∞<br />

r j


4.3 A BIPOLAR POWER SPECTRUM STUDY OF COSMIC TOPOLOGY 42<br />

one could use the bipolar power spectrum as a probe to detect the topology of the<br />

universe.<br />

When <strong>CMB</strong> anisotropy is multiply imaged, the bipolar power spectrum corresponds<br />

to a correlation pattern of matched pairs of circles. Since it is orientation independent,<br />

the BiPS has to be computed only once for the <strong>CMB</strong> map irrespective of the orientation<br />

of the DD <strong>with</strong> respect to the sky.<br />

This method is very fast because there are fast<br />

spherical harmonic transform methods of the <strong>CMB</strong> sky maps.<br />

Using these methods<br />

it is very fast and efficient to compute the BiPS even for maps in WMAP resolution<br />

(HEALPix, N side = 512) in a few minutes on a single processor. We have used BiPS to<br />

put constraints on the topology of the Universe, Hajian et al. (2006).<br />

The A lM<br />

l 1 l 2<br />

signature, which is not discussed here, contains more details of orientation<br />

of the SI violation. It may be possible to detect the A lM<br />

l 1 l 2<br />

too for strong violation of SI.<br />

In what follows we use the method of images to compute the two point correlations<br />

in compact spaces. BiPS is then computed from two point correlations. What we are<br />

interested in is the signature of compact spaces on BiPS and its measurability from a<br />

single map. We study two families of multi-connected spaces which are presented below.<br />

4.3.1 Flat Spaces<br />

Among all multi-connected three-dimensional spaces, “flat spaces” have been studied<br />

the most extensively in the cosmological context. This is because of the computational<br />

simplicity of these spaces. There are 18 multi-connected flat spaces, given as the quotient<br />

E 3 /Γ of three-dimensional Euclidean space E 3 by a group Γ of symmetries of E 3<br />

that is discrete and fixed point free. The classification of these spaces has long been<br />

known. Among these 18 spaces, 10 are quotients of the 3-torus, 6 of them are orientable<br />

and 4 are non-orientable. Three of these spaces have a cubic fundamental domain, 6<br />

have hexagonal fundamental domains and one of them has a rhombic dodecahedron circumscribed<br />

about a rectangular box as its fundamental polyhedron. These fundamental<br />

polyhedra are shown in Fig. 4.1. These spaces are compact in three dimensions. There<br />

are two partially compact families as well; the chimney space and its quotients which


4.3 A BIPOLAR POWER SPECTRUM STUDY OF COSMIC TOPOLOGY 43<br />

Symbol Name NC Volume Orientable<br />

E 1 3-torus 3 Finite Yes<br />

E 2 half turn space 3 Finite Yes<br />

E 3 quarter turn space 3 Finite Yes<br />

E 4 third turn space 3 Finite Yes<br />

E 5 sixth turn space 3 Finite Yes<br />

E 6 Hantzsche-Wendt space 3 Finite Yes<br />

E 7 Klein space 3 Finite No<br />

E 8 Klein space <strong>with</strong> horizontal flip 3 Finite No<br />

E 9 Klein space <strong>with</strong> vertical flip 3 Finite No<br />

E 10 Klein space <strong>with</strong> half turn 3 Finite No<br />

E 11 chimney space 2 ∞ Yes<br />

E 12 chimney space <strong>with</strong> half turn 2 ∞ Yes<br />

E 13 chimney space <strong>with</strong> vertical flip 2 ∞ No<br />

E 14 chimney space <strong>with</strong> horizontal flip 2 ∞ No<br />

E 15 chimney space <strong>with</strong> half turn and flip 2 ∞ No<br />

E 16 slab space 1 ∞ Yes<br />

E 17 slab space <strong>with</strong> flip 1 ∞ No<br />

E 18 Euclidean space 0 ∞ Yes<br />

Table 4.1 Classification of the 18 three-dimensional flat spaces (NC means number of<br />

compact dimensions). Table reproduced from Riazuelo et al.<br />

(2004b).


4.3 A BIPOLAR POWER SPECTRUM STUDY OF COSMIC TOPOLOGY 44<br />

3-Torus Quarter Turn Space Half Turn Space<br />

Sixth Turn Space Third Turn Space Hantzsche-Wendt Space<br />

Klein Space<br />

Klein Space<br />

<strong>with</strong> Horizontal Flip<br />

Klein Space<br />

<strong>with</strong> Vertical Flip<br />

Klein Space<br />

<strong>with</strong> Half Turn<br />

Figure 4.1 Fundamental domains for the compact flat three-manifolds. The unmarked<br />

walls are glued straight across. Figure from Riazuelo et al.<br />

(2004b).<br />

have two compact dimensions (fundamental polyhedra are shown in Fig. 4.2), and the<br />

slab space <strong>with</strong> its quotient <strong>with</strong> only one compact dimension (figure 4.3). Summary of<br />

properties of these spaces is given in table 4.1. A detailed study on the signature of flat<br />

spaces on <strong>CMB</strong> has been done by Riazuelo et al. (2004b).<br />

The simple torus, T 3 , is the simplest compact space and one can in fact calculate<br />

most of its properties analytically. Here we will study this simple case which indeed has<br />

very interesting properties.


4.3 A BIPOLAR POWER SPECTRUM STUDY OF COSMIC TOPOLOGY 45<br />

Chimney Space<br />

Chimney Space <strong>with</strong><br />

Half Turn<br />

Chimney Space <strong>with</strong><br />

Vertical Flip<br />

Chimney Space <strong>with</strong><br />

Horizontal Flip<br />

Chimney Space <strong>with</strong><br />

Half Turn and Flip<br />

Figure 4.2 The chimney space is made from an infinitely tall rectangular chimney <strong>with</strong><br />

front and back faces (resp. left and right faces) glued straight across. The four variations<br />

on the chimney space space glue the front face to the back as indicated by the doors. In<br />

all variations except the last the left and right faces are glued straight across; in the last<br />

variation they are glued <strong>with</strong> a top-to-bottom flip so that the windows match. Figure<br />

from Riazuelo et al. (2004b).<br />

Slab Space<br />

Slab Space <strong>with</strong> Flip<br />

Figure 4.3 The slab space is made from an infinitely tall and wide slab of space <strong>with</strong><br />

its front face glued to its back face straight across. The variation glues the faces <strong>with</strong> a<br />

flip. Figure from Riazuelo et al. (2004b).


4.3 A BIPOLAR POWER SPECTRUM STUDY OF COSMIC TOPOLOGY 46<br />

Two point correlation<br />

In a simple T 3 space which is defined by identifying x → x + L, y → y + L and<br />

z → z + L, the eigenvalue spectrum is discretized as<br />

<strong>with</strong><br />

k = 2π L (n 1ˆx + n 2 ŷ + n 3 ẑ) = 2π n = kˆn, (4.32)<br />

L<br />

n ≡ (n 1 , n 2 , n 3 ), (4.33)<br />

n ≡ √ n · n, (4.34)<br />

ˆn ≡ (n 1 , n 2 , n 3 )/n, (4.35)<br />

k ≡ √ k · k = 2π n, (4.36)<br />

L<br />

Considering only naive Sachs-Wolfe effect, the correlation function will become<br />

ξ [Γ] (x, x ′ ) =<br />

1 L 3 ∑<br />

n<br />

P φ (k n ) exp (−i 2πn<br />

L · (x − x′ )). (4.37)<br />

The above expression is in fact a cosine transform of the primordial power spectrum<br />

P (k). To see this, we should break the summation over n into summation over positive<br />

and negative n. And the one can use the definition cos x = (exp (ix) + exp (−ix))/2 to<br />

obtain<br />

ξ [Γ] (x, x ′ ) = 23<br />

L 3<br />

= 23<br />

L 3<br />

= 23<br />

L 3<br />

= 23<br />

L 3<br />

∑<br />

n x,n y,n z>0<br />

∑<br />

∑<br />

n x>0 n y>0 n z>0<br />

∑<br />

∑<br />

P φ (k n ) cos 2πn x∆x<br />

L<br />

∑<br />

∑<br />

n x>0 n y>0 n z>0<br />

∑<br />

∑<br />

∑<br />

n x>0 n y>0 n z>0<br />

cos 2πn y∆y<br />

L<br />

cos 2πn z∆z<br />

L<br />

k 3<br />

k P φ(k 3 n ) cos 2πn x∆x<br />

cos 2πn y∆y<br />

cos 2πn z∆z<br />

L L L<br />

P φ (k n )<br />

k 3<br />

cos 2πn x∆x<br />

L<br />

cos 2πn y∆y<br />

L<br />

cos 2πn z∆z<br />

L<br />

( L<br />

2πn )3 P φ (k n ) cos 2πn x∆x<br />

cos 2πn y∆y<br />

L L<br />

cos 2πn z∆z<br />

(4.38) .<br />

L<br />

This is nothing but discrete Fourier transform of P φ (k n ). One can use FFT methods<br />

to compute the two point correlations in T 3 spaces. A 2-dimensional example of this is<br />

shown in figure 4.4.<br />

In an Euclidean space when the separation between two points increases, the correlation<br />

between them decreases. For a compact space it is different. We know that all


4.3 A BIPOLAR POWER SPECTRUM STUDY OF COSMIC TOPOLOGY 47<br />

correlation function using 3D FFT analysis (computed for 4 Dirichlet Domains)<br />

6<br />

4<br />

2<br />

0<br />

−2<br />

140<br />

120<br />

100<br />

80<br />

100<br />

120<br />

140<br />

60<br />

80<br />

40<br />

20<br />

0<br />

0<br />

20<br />

40<br />

60<br />

Figure 4.4 Two point correlations in a 2-dimensional Toroidal space of the size of 64<br />

units. The correlation is between the central point and other points on the surface and<br />

is computed using the FFT method. Figure from Kourkchi (2006).<br />

the light received from a specific time, e.g. the epoch of the last scattering, represents a<br />

sphere. But if this sphere is larger than the universe it will intersect itself. If the space<br />

is a torus, then the sphere will intersect the imaginary walls of a “periodic” universe,<br />

and hence will intersect itself. This intersection has a very important effect on the large<br />

scale properties of the sphere for it identifies parts of the sphere which are apparently<br />

very far from each other. This will show up in the two point correlation as unusually<br />

high correlation between widely separate points on the sky.<br />

Consider a great circle on the surface of last scattering. This circle is greater than<br />

the size of the fundamental domain so it will fold back on itself and make a pattern like<br />

what is shown in Fig.4.5. Intersection of surface of last scattering <strong>with</strong> itself in small<br />

universes defines identified circles in the sky (as it is shown in Fig. 4.7). Search for


4.3 A BIPOLAR POWER SPECTRUM STUDY OF COSMIC TOPOLOGY 48<br />

4<br />

3<br />

2<br />

1<br />

y<br />

0<br />

-1<br />

-2<br />

-3<br />

-4<br />

-4 -3 -2 -1 0 1 2 3 4<br />

x<br />

Figure 4.5 A circle <strong>with</strong> a radius greater than the size of the fundamental domain will<br />

fold back on itself.<br />

1.1<br />

1<br />

0.9<br />

0.8<br />

Correlation<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0 0.5 1 1.5 2<br />

Angular Separation/Pi<br />

Figure 4.6 Correlation of two points on a circle whose radius is equal to the size of the<br />

fundamental domain. Strong correlations between apparently far apart points are seen.


4.3 A BIPOLAR POWER SPECTRUM STUDY OF COSMIC TOPOLOGY 49<br />

these identified circles has been going on in the last few years by Cornish, Spergel &<br />

Starkman (1998); Cornish et al. (2003) and is a nice and direct method to find the<br />

topology of the universe. We use the method of images Bond, Pogosyan & Souradeep<br />

Figure 4.7 Identified circles in the sky along the x, y and z axes, for a toroidal space<br />

<strong>with</strong> R SLS = 0.65L.<br />

(1998, 2000) to compute the two point correlations on this circle as a function of the<br />

angular separation between them. The result is interesting: correlation falls off and<br />

again rises as we increase the angular separation between the two points. This result<br />

is shown in Fig.4.6. If we choose another great circle on the sphere, the shape of the<br />

correlation function and its dependence to θ would be completely different because it<br />

depends on the way the circle has folded back on itself. This shows that in flat torus<br />

the statistical isotropy breaks down.


NONP<br />

4.3 A BIPOLAR POWER SPECTRUM STUDY OF COSMIC TOPOLOGY 50<br />

ο<br />

α=75<br />

α=45 ο<br />

α=60 ο QRQ<br />

SRS<br />

Figure 4.8 The pattern in <strong>CMB</strong> correlation in a multi-connected universe, M [Γ] ≡<br />

M [X] /Γ is related to the distribution of ‘images’ of the sphere of last scattering (SLS) on<br />

the universal cover. The three figures illustrate this for three values of α = 45 ◦ , 60 ◦ , 75 ◦<br />

in a squeezed torus. The solid parallelepiped is the fundamental domain and the dashed<br />

show its images tessellate,M [X] . The solid circle is the SLS and dashed/dotted ones are<br />

its images. In each case, the radius of SLS is equal to the in-radius, R < of M [Γ] . The<br />

shaded polygon is the Dirichlet domain (DD). The closest SLS images which determine<br />

the DD are dashed (others are dotted). Note the hexagonal shape of DD when α ≠ π/2<br />

and is equal sided for α = π/3. For smaller α, the DD approximates an elongated<br />

cuboid.<br />

BiPS of Toroidal Spaces<br />

We restrict attention to the case where <strong>CMB</strong> anisotropy arises entirely at the sphere<br />

of last scattering (SLS) of radius R ∗ (nearly equal to observable horizon).<br />

Invoking<br />

method of images, the <strong>CMB</strong> correlation pattern on the SLS is known to be dictated by<br />

the distribution of nearest ‘images’ of the SLS on the universal cover Bond, Pogosyan<br />

& Souradeep (1998, 2000). The correlations are distorted even when the SLS and its<br />

images do not intersect (R ∗ < R < ). When SLS intersects its images the <strong>CMB</strong> sky is<br />

multiply imaged in characteristic correlation pattern of pairs of circles Cornish, Spergel<br />

& Starkman (1998). Fig. 4.8 illustrates the role of the nearest SLS images and related<br />

DD in defining the principal directions in the correlation for a few cases.<br />

We compute C(ˆq, ˆq ′ ) for <strong>CMB</strong> anisotropy in torus space using regularized method<br />

of images Bond, Pogosyan & Souradeep (1998, 2000). We can compute the κ l in real<br />

space using eq. (2.30) or from A lM<br />

ll ′<br />

using eq. (2.20) in harmonic space. Fig. 4.9 plots<br />

the κ l spectrum for a number of cubic, cuboidal and squeezed torus spaces. We find the<br />

following interesting results :


4.3 A BIPOLAR POWER SPECTRUM STUDY OF COSMIC TOPOLOGY 51<br />

2<br />

1.5<br />

1<br />

0.5<br />

Cubic<br />

(big)<br />

R=0.4L<br />

R=0.5L<br />

R=0.6L<br />

0<br />

1 2 3 4 5 6 7 8 9<br />

8<br />

L 2<br />

=5 L<br />

Cuboid<br />

1<br />

6<br />

eql. vol.<br />

50 Cubic R=1.0L<br />

(small) R=1.5L<br />

40<br />

R=2.0L<br />

30<br />

20<br />

10<br />

0<br />

1 2 3 4 5 6 7 8 9 10 11 12 13<br />

4<br />

L 2<br />

=5 L<br />

Cuboid<br />

1<br />

3<br />

R=R L 2<br />

=2 L <<br />

1<br />

L 2<br />

=2 L 1<br />

1 2 3 4 5 6 7 8 9<br />

4<br />

2<br />

2<br />

1<br />

0<br />

1 2 3 4 5 6 7 8 9<br />

0<br />

0.8<br />

2<br />

0.6<br />

Squeezed<br />

45 Deg<br />

eql. vol.<br />

60 Deg<br />

75 Deg<br />

1.5<br />

0.4<br />

1<br />

Squeezed<br />

R=R <<br />

45 Deg<br />

60 Deg<br />

75 Deg<br />

0.2<br />

0<br />

1 2 3 4 5 6 7 8 9<br />

0.5<br />

0<br />

1 2 3 4 5 6 7 8 9<br />

Figure 4.9 The κ l spectra for flat tori models are plotted. The top row panels are for<br />

cubic tori spaces. The left panel shows spaces of volume, V M [Γ], larger than the volume V ∗<br />

contained in the sphere of last scattering (SLS) <strong>with</strong> V M [Γ]/V ∗ = 3.7, 1.9, 1.1, respectively.<br />

The right panel shows small spaces <strong>with</strong> V M [Γ]/V ∗ = 0.24, 0.07, 0.03, respectively. Note<br />

that κ 2 = 0 for cubic tori. The middle panels consider cuboid tori <strong>with</strong> 1 : 5 and<br />

1 : 2 ratio of identification lengths. The bottom panels show κ l for equal-sided squeezed<br />

tori <strong>with</strong> α = 45 ◦ , 60 ◦ and 75 ◦ . In the middle and bottom rows, the right panels show<br />

the case when radius of SLS, R ∗ = R < the in-radius of the space. Here, the SLS just<br />

touches its nearest images (see Fig. 4.8) which is at the threshold where <strong>CMB</strong> anisotropy<br />

is multiply imaged for larger R ∗ . The cases in the left panels of lower two rows have<br />

V M [Γ]/V ∗ = 1 and are at the divide between large and small spaces.


4.3 A BIPOLAR POWER SPECTRUM STUDY OF COSMIC TOPOLOGY 52<br />

i. κ l = 0 for odd l for all torus models. This feature has been studied by Mukhopadhyay<br />

(2005) and can be explained by 2n-fold rotational symmetries of the space about<br />

an axis. This does not hold for compact hyperbolic spaces.<br />

ii. For cubic torus κ 2 = 0. κ 2 is non-zero for cuboidal and squeezed torus. This is a<br />

clear signature of non-cubic torus where the DD differs from the FD and has more than<br />

three principle axes.<br />

iii. For equal-sided squeezed torus, κ 4 , decreases as α decreases from 90 ◦ to 60 ◦<br />

as R < increases. For α < 60 ◦ sharply increases <strong>with</strong> decreasing α as R < decreases<br />

sharply Bowen & Ferreira (2002).<br />

iv. κ 2 = 0 increases monotonically as α decreases from 90 ◦ . The trend is well fit by<br />

eq. (4.88).<br />

v. The peak of κ l shifts to larger l for small spaces.<br />

The results can be understood using the leading order terms of the correlation function<br />

in a torus where κ l can be calculated analytically. The spatial correlation function<br />

of gravitational potential in the periodic box implied by the T 3 topology <strong>with</strong> cubic<br />

fundamental domain is given by eqn. (4.37)<br />

ξ Φ (x, x ′ ) = L −3 ∑ n<br />

2πn<br />

−i<br />

P Φ (k n ) e L ·(x−x′) , (4.39)<br />

where n is 3-tuple of integers, k 2 n = (2π/L) 2 (n · n), and the term <strong>with</strong> n · n = 0 is<br />

excluded from the summation. We restrict attention to the case where <strong>CMB</strong> anisotropy<br />

arises entirely at the sphere of last scattering (SLS) of radius R SLS . Two point correlation<br />

function of <strong>CMB</strong> anisotropy is obtained from the spatial correlation function of<br />

gravitational potential on the SLS along the two directions ˆq 1 and ˆq 2 as<br />

C T 3 (ˆq 1 , ˆq 2 ) = ξ T 3<br />

Φ (R SLS ˆq 1 , R SLS ˆq 2 ) (4.40)<br />

= 1 ∑<br />

P<br />

L 3 Φ (kˆn ) exp −iπR SLS<br />

(ˆn · ˆq 1 − ˆn · ˆq 2 )<br />

L/2<br />

ˆn<br />

Figure 4.10 shows the relative configuration of ˆq 1 , ˆq 2 and ∆ˆq. Using the simple geometry<br />

of this configuration, we can find the vector subtraction which is needed to compute the<br />

above two point correlation; Using cosine rules in triangles, R SLS ∆ˆq is given by<br />

R SLS |∆ˆq| = R SLS<br />

√<br />

2 − 2 cos θ = 2RSLS sin θ/2 (4.41)<br />

=⇒ R SLS (ˆq 1 − ˆq 2 ) = 2R SLS sin θ/2∆ˆq


4.3 A BIPOLAR POWER SPECTRUM STUDY OF COSMIC TOPOLOGY 53<br />

R<br />

SLS ∆ q^<br />

R SLS<br />

q^ 2<br />

θ<br />

R SLS<br />

q^ 1<br />

Figure 4.10 Geometry of the distance between two points along the two directions ˆq 1<br />

and ˆq 2 on the surface of last scattering; R SLS ∆ˆq = 2R SLS sin θ/2∆ˆq.<br />

Substituting this into the expression for two point correlation in toroidal spaces, eqn.(4.40),<br />

we get<br />

C T 3 (ˆq 1 , ˆq 2 ) =<br />

1 ∑<br />

P<br />

L 3 Φ (kˆn )e −2iπ P i ɛ in i ∆q i<br />

(4.42)<br />

ˆn<br />

where the small parameter ɛˆq ≤ 1 is the physical distance to the SLS along ˆq in units of<br />

L i /2, where L i are sizes of three dimensions of the fundamental domain.<br />

ɛ i = 2R SLS<br />

L i<br />

sin θ/2 θ = cos −1 ˆq 1 · ˆq 2 (4.43)<br />

Contribution to C(ˆq, ˆq ′ ) from large wavenumbers, |n|ɛ ≫ 1 is expected to be approximately<br />

statistically isotropic. The SI violation is found in the low wavenumbers.<br />

When the SLS is contained <strong>with</strong> the fundamental domain around the observer, i.e.,<br />

the <strong>CMB</strong> anisotropy is not multiply imaged on the sky, 2R SLS /L is a constant. When<br />

ɛ is a small constant, the leading order terms in the correlation function eq. (4.42) can<br />

be readily obtained in power series expansion in powers of ɛ. In what follows, we will<br />

compute the leading order terms in the correlation function in powers of ɛ for the lowest<br />

wavenumbers |n| 2 = 1, |n| 2 = 2 and |n| 2 = 3 in a cubic FD torus.<br />

ˆn · ˆn = 1<br />

The lowest wavenumbers are |n| 2 = 1 which means<br />

n x = ±1; n y = 0; n z = 0 (4.44)


4.3 A BIPOLAR POWER SPECTRUM STUDY OF COSMIC TOPOLOGY 54<br />

<strong>with</strong> a cyclic relation between n x , n y and n z . The correlation function in a cubic torus<br />

space for these wavenumbers will be<br />

C (1) (ˆq 1 , ˆq 2 ) = 1 L 3 P Φ( 2π L ) [ exp 2iπɛ 1∆q 1 + exp −2iπɛ 1 ∆q 1 (4.45)<br />

This can be expanded in powers of ɛ i ,<br />

+ exp 2iπɛ 2 ∆q 2 + exp −2iπɛ 2 ∆q 2<br />

+ exp 2iπɛ 3 ∆q 3 + exp −2iπɛ 3 ∆q 3 ]<br />

= 2 L P Φ( 2π 3∑<br />

3 L ) cos 2πɛ i ∆q i<br />

i=1<br />

C (1) (ˆq 1 , ˆq 2 ) ≈<br />

2 L 3 P Φ( 2π L<br />

)[3 −<br />

(2π)2<br />

2!<br />

3∑<br />

i=1<br />

ɛ 2 i ∆q 2 i + (2π)4<br />

4!<br />

3∑<br />

ɛ 4 i ∆qi 4 ]. (4.46)<br />

i=1<br />

The above expression is for an arbitrary Cuboid torus <strong>with</strong> un-equal sides. In a cubic<br />

torus, all three sides are equal, i.e.<br />

L i = L =⇒ ɛ 3 = ɛ 2 = ɛ 1 (4.47)<br />

And hence the leading order terms in the correlation function will be<br />

C 1 (ˆq 1 , ˆq 2 ) = C 0 [1 − ɛ2<br />

2!<br />

3∑<br />

i=1<br />

∆q 2 i + 3ɛ4<br />

4!<br />

3∑<br />

∆qi 4 ] (4.48)<br />

i=1<br />

Where<br />

C 0 = 6 L 3 P Φ( 2π L ) (4.49)<br />

ɛ = 2πɛ 1<br />

√<br />

3<br />

.<br />

This correlation function can be written as a superposition of toy models we have studied<br />

so far. The second term is toy model I <strong>with</strong> three axis and third is the fourth order toy<br />

model, so we have<br />

C (1)<br />

T 3 = Toy Model I With Three Axes ( O(ɛ 2 ) ) + Fourth Order Toy Model ( O(ɛ 4 ) )<br />

We retain O(ɛ 4 ) terms because as we have seen, Toy Model I up to the second order<br />

in ɛ, O(ɛ 2 ), is rotationally invariant and hence does not contribute to the violation of


4.3 A BIPOLAR POWER SPECTRUM STUDY OF COSMIC TOPOLOGY 55<br />

SI. From the above estimations we should expect a non-zero κ 4 in BiPS. The non-zero<br />

κ l can be analytically computed to be<br />

κ l =<br />

(2l + 1)2<br />

(8π 2 ) 2 [κ 0 δ l0 + κ 4 δ l4 ] , (4.50)<br />

where<br />

κ 0<br />

C0<br />

2<br />

κ 4<br />

C0<br />

2<br />

= π 2 (1 − 4ɛ 2 + 368<br />

15 ɛ4 − 288<br />

5 ɛ6 + 20736<br />

125 ɛ8 )<br />

= 12288π2 ɛ 8 (4.51)<br />

875<br />

As we expected, the first non zero κ l in cubic (equal-sided) torus occurs at l = 4. κ 4 ∼ ɛ 8<br />

falls off rapidly as ɛ → 0 for large spaces.<br />

ˆn · ˆn = 2<br />

The next lowest wavenumbers are those of |n| 2 = 2, which can be like<br />

n x = n y = ±1; n z = 0 (4.52)<br />

and any cyclic variant of them. The correlation function in a cubic torus space for these<br />

wavenumbers will be<br />

C (2)<br />

T 3 (ˆq 1 , ˆq 2 ) =<br />

1 L P Φ( 2√ 2π<br />

3 L ) [ exp −2iπ(ɛ 1∆q 1 + ɛ 2 ∆q 2 ) (4.53)<br />

+ exp −2iπ(ɛ 1 ∆q 1 − ɛ 2 ∆q 2 )<br />

+ exp −2iπ(−ɛ 1 ∆q 1 + ɛ 2 ∆q 2 )<br />

+ exp −2iπ(−ɛ 1 ∆q 1 − ɛ 2 ∆q 2 )<br />

+ similar terms for ɛ 2 & ɛ 3 ]<br />

= 2 L P Φ( 2√ 2π<br />

3 L ) [ cos (2πɛ 1∆q 1 + 2πɛ 2 ∆q 2 ) + cos (2πɛ 1 ∆q 1 − 2πɛ 2 ∆q 2 )<br />

+ cos (2πɛ 1 ∆q 1 + 2πɛ 3 ∆q 3 ) + cos (2πɛ 1 ∆q 1 − 2πɛ 3 ∆q 3 )<br />

+ cos (2πɛ 2 ∆q 2 + 2πɛ 3 ∆q 3 ) + cos (2πɛ 2 ∆q 2 − 2πɛ 3 ∆q 3 ) ].<br />

The above expressions can be simplified if we use the following rule of cosines<br />

cos (α + β) + cos (α − β) = 2 cos α cos β, (4.54)


4.3 A BIPOLAR POWER SPECTRUM STUDY OF COSMIC TOPOLOGY 56<br />

We can then expand it in powers of ɛ i ,<br />

C (2)<br />

T 3 (ˆq 1 , ˆq 2 ) =<br />

4 L P Φ( 2√ 2π<br />

3 L ) [ cos 2πɛ 1∆q 1 cos 2πɛ 2 ∆q 2 (4.55)<br />

+ cos 2πɛ 1 ∆q 1 cos 2πɛ 3 ∆q 3<br />

≈ 4 L 3 P Φ( 2√ 2π<br />

L<br />

= 4 L 3 P Φ( 2√ 2π<br />

+ cos 2πɛ 2 ∆q 2 cos 2πɛ 3 ∆q 3 ]<br />

) [ (1 −<br />

(2π)2<br />

2!<br />

(1 − (2π)2<br />

2!<br />

+ (1 − (2π)2<br />

2!<br />

(1 − (2π)2<br />

2!<br />

+ (1 − (2π)2<br />

2!<br />

(1 − (2π)2<br />

2!<br />

L ) [ 3 − (2π)2 3∑<br />

+ (2π)4<br />

4!<br />

ɛ 2 1 ∆q2 1 + (2π)4 ɛ 4 1<br />

4!<br />

∆q4 1 )<br />

ɛ 2 2∆q2 2 + (2π)4 ɛ 4<br />

4!<br />

2∆q2)<br />

4<br />

ɛ 2 1∆q1 2 + (2π)4 ɛ 4<br />

4!<br />

1∆q1)<br />

4<br />

ɛ 2 3∆q3 2 + (2π)4 ɛ 4<br />

4!<br />

3∆q3)<br />

4<br />

ɛ 2 2∆q2 2 + (2π)4 ɛ 4<br />

4!<br />

2∆q2)<br />

4<br />

ɛ 2 3 ∆q2 3 + (2π)4 ɛ 4 3<br />

4!<br />

∆q4 3 ) ]<br />

3∑<br />

i=1<br />

ɛ 2 i ∆q2 i + 2(2π)4<br />

4!<br />

∑<br />

ɛ 2 i ɛ2 j ∆q2 i ∆q2 j ]<br />

i,j<br />

i=1<br />

ɛ 4 i ∆q4 i<br />

And finally the two point correlation is<br />

[<br />

3∑<br />

C (2)<br />

T<br />

(ˆq 3 1 , ˆq 2 ) ≈ C 0 1 − ɛ 2 ∆qi 2 + 3 2 ɛ4<br />

in which<br />

i=1<br />

3∑<br />

i=1<br />

∆q 4 i + 9 4 ɛ4 ∑ i,j<br />

∆q 2 i ∆q2 j<br />

]<br />

. (4.56)<br />

ɛ 2 = 8(2π) 2 ɛ 2 1 (4.57)<br />

C 0 =<br />

1<br />

18L P Φ( 2√ 2π<br />

3 L ).<br />

The above correlation function can be written as a superposition of toy models we have<br />

studied so far.<br />

C (2)<br />

T 3 = Toy Model I With Three Equal Axes ( O(ɛ 2 ) ) (4.58)<br />

+ Fourth Order Toy Model ( O(ɛ 4 ) )<br />

+ Second Order Cross Terms ( O(ɛ 4 ) )


4.3 A BIPOLAR POWER SPECTRUM STUDY OF COSMIC TOPOLOGY 57<br />

Among the toy models on the RHS of the above expression, the first one is rotationally<br />

invariant and does not contribute to non-trivial components (l ≠ 0) of BiPS. The second<br />

term as well as the third term will cause a non-zero κ 4 which will be of the order of<br />

O(ɛ 8 ).<br />

ˆn · ˆn = 3<br />

κ l =<br />

(2l + 1)2<br />

(8π 2 ) 2 [κ 0 δ l0 + κ 4<br />

(<br />

O(ɛ 8 ) ) δ l4 ] , (4.59)<br />

The next lowest wavenumbers are those of |n| 2 = 2, which are<br />

n x = ±1; n y = ±1; n z = ±1. (4.60)<br />

Substituting these in the expression for the correlation function, we see that the correlation<br />

function in a cubic torus space for these wavenumbers will again have a cosine<br />

structure<br />

C (3)<br />

T 3 (ˆq 1 , ˆq 2 ) =<br />

4 L P Φ( 2√ 2π<br />

3 L ) [ cos 2π(ɛ 1∆q 1 + ɛ 2 ∆q 2 + ɛ 3 ∆q 3 ) (4.61)<br />

+ cos 2π(ɛ 1 ∆q 1 − ɛ 2 ∆q 2 + ɛ 3 ∆q 3 )<br />

+ cos 2π(ɛ 1 ∆q 1 + ɛ 2 ∆q 2 − ɛ 3 ∆q 3 )<br />

+ cos 2π(−ɛ 1 ∆q 1 + ɛ 2 ∆q 2 + ɛ 3 ∆q 3 ) ]<br />

= 12<br />

L 3 P Φ( 2√ 2π<br />

L ) [ cos 2πɛ 1∆q 1 cos 2πɛ 2 ∆q 2 cos 2πɛ 3 ∆q 3 ].<br />

Expanding the above expression in powers of ɛ i , we will obtain<br />

C (3)<br />

T 3 (ˆq 1 , ˆq 2 ) ≈ 12<br />

L P Φ( 2√ 2π<br />

3 L<br />

) [ (1 −<br />

(2π)2<br />

2!<br />

× (1 − (2π)2<br />

2!<br />

ɛ 2 1 ∆q2 1 + (2π)4 ɛ 4 1<br />

4!<br />

∆q4 1 ) (4.62)<br />

ɛ 2 2 ∆q2 2 + (2π)4 ɛ 4 2<br />

4!<br />

∆q4 2 )<br />

× (1 − (2π)2 ɛ 2 2<br />

2!<br />

∆q2 2 + (2π)4 ɛ 4 2<br />

4!<br />

∆q4 2 ) ]<br />

= 12<br />

L P Φ( 2√ 2π<br />

3∑<br />

3∑<br />

3 L ) [ 3 − (2π)2 ɛ 2 i<br />

2!<br />

∆q2 i + (2π)4 ɛ 4 i<br />

4!<br />

∆q4 i<br />

+ (2π)4<br />

4!<br />

i=1<br />

∑<br />

ɛ 2 i ɛ 2 j∆qi 2 ∆qj 2 ]<br />

For a cubic (equal-sided) torus, we have ɛ i = ɛ. And the two point correlation will be<br />

3∑<br />

3∑<br />

3∑<br />

C (3)<br />

T<br />

(ˆq 3 1 , ˆq 2 ) ≈ C 0 [1 − ɛ 2 ∆qi 2 + ɛ4 ∆qi 4 4<br />

+ ɛ4 ∆qi 4 4<br />

q4 j ]. (4.63)<br />

i=1<br />

i=1<br />

i,j<br />

i≠j<br />

i=1


4.3 A BIPOLAR POWER SPECTRUM STUDY OF COSMIC TOPOLOGY 58<br />

in which<br />

ɛ 2 = (2π)2<br />

3 ɛ2 1 (4.64)<br />

C 0 = 18<br />

L 3 P Φ( 2√ 2π<br />

L ).<br />

The above correlation function can be written as a superposition of toy models we have<br />

studied so far.<br />

C (3)<br />

T 3 = Toy Model I With Three Equal Axes ( O(ɛ 2 ) ) (4.65)<br />

+ Fourth Order Toy Model ( O(ɛ 4 ) )<br />

+ Second Order Cross Terms ( O(ɛ 4 ) )<br />

Among the toy models on the RHS of the above expression, the first one is rotationally<br />

invariant and does not contribute to non-trivial components (l ≠ 0) of BiPS. The second<br />

term as well as the third term will cause a non-zero κ 4 which will be of the order of<br />

O(ɛ 8 ).<br />

Cuboid Torus<br />

κ l =<br />

(2l + 1)2<br />

(8π 2 ) 2 [κ 0 δ l0 + κ 4<br />

(<br />

O(ɛ 8 ) ) δ l4 ] , (4.66)<br />

Till now we have studied the equal sided torus spaces in which the dimensions of the<br />

fundamental domain are the same and as a result, ɛ i = ɛ. A more general case is cuboidal<br />

(unequal-sided) torus in which ɛ i are different. This can be constructed from the equal<br />

sided torus by the act of a rescaling operator, ρ. This operator is a diagonal 3×3 matrix<br />

which re-scales the sizes of the fundamental domain by different identification scales in<br />

each direction. This matrix is given by<br />

ρ =<br />

⎛<br />

⎞<br />

R 1 0 0<br />

⎜<br />

⎝ 0 R 2 0 ⎟<br />

⎠ (4.67)<br />

0 0 R 3<br />

This transformation in the first place changes the volume of the fundamental domain<br />

by a factor of R 1 R 2 R 3 , and on the other hand it induces anisotropic scale differences in


4.3 A BIPOLAR POWER SPECTRUM STUDY OF COSMIC TOPOLOGY 59<br />

the sides of the fundamental domain and changes it into a rectangular parallelepiped.<br />

In that case for wavenumbers of |n| 2 = 1, the correlation violates SI at order ɛ 2 and to<br />

that order is given as 2 C 1 (ˆq 1 , ˆq 2 ) = C 0<br />

[<br />

1 − ɛ2<br />

2!<br />

]<br />

3∑<br />

βi 2 ∆qi<br />

2<br />

i=1<br />

(4.68)<br />

in which ɛ is the physical distance to the SLS along ˆq in units of ¯L/2, where ¯L is the<br />

geometrical average of sizes of three dimensions of the fundamental domain, L i . And β i<br />

is the ratio of size of each dimension and the average size,<br />

ɛ 2 = (2π)2<br />

3 (R SLS<br />

¯L )2 (2 sin θ/2) 2 (4.69)<br />

β 2 i = ( ¯L<br />

L i<br />

) 2 .<br />

The above expression is the one for Toy Model I, <strong>with</strong> three different axes,<br />

C (3)<br />

T 3 = Toy Model I With Three Un − Equal Axes ( O(ɛ 2 ) )<br />

and hence, should have a non-zero κ 2 of the order of O(ɛ 4 ).<br />

The non zero κ l corresponding to the correlation eqn. (4.68) can be analytically<br />

calculated to be<br />

κ 0<br />

C0<br />

2<br />

κ 2<br />

C0<br />

2<br />

= π 2 (1 − 4 3 β2 ɛ 2 + 16<br />

27 β4 ɛ 4 )<br />

= 64π2<br />

135<br />

[<br />

β 4 − 3(β 2 1β 2 2 + β 2 1β 2 3 + β 2 2β 2 3) ] ɛ 4 (4.70)<br />

where β 2 = ∑ 3<br />

i=1 β2 i . The κ 2 ∼ ɛ 4 signal falls off slower than the κ 4 signal in cubic space<br />

for large spaces (ɛ → 0).<br />

Squeezed Torus<br />

Next, we study the flat toroidal spaces in their most general form. We studied the<br />

cases of unequal-sided orthogonal cubes as the fundamental domain which were obtained<br />

2 This is in fact the |n| 2 = 1 result for cubical torus <strong>with</strong> un-equal ɛ i


4.3 A BIPOLAR POWER SPECTRUM STUDY OF COSMIC TOPOLOGY 60<br />

from the simple cubic torus by linear transformations that did not keep the volume<br />

conserved, but kept the cube orthogonal. We can now allow for linear transformations<br />

that change the angles of the corner of the fundamental domain while preserving its<br />

volume. This can be done by a shape transformation matrix, S,<br />

S = ω<br />

⎛<br />

⎞<br />

1 cos α 1 cos α 2<br />

⎜<br />

⎝ 0 sin α 1 cos α 3 sin α 2<br />

⎟<br />

⎠ (4.71)<br />

0 0 sin α 2 sin α 3<br />

where ω is defined as<br />

ω = (sin α 1 sin α 2 sin α 3 ) 1/3 (4.72)<br />

and ensures the unity of the shape transformation matrix detS = 1.<br />

If we apply this transformation on a unit cube defined by diag(1, 1, 1), we will get a<br />

unit parallelepiped in the following form<br />

S<br />

ˆx ŷ ẑ ˆx S ŷ S ẑ S<br />

⎛ ⎞ ⎛<br />

⎞<br />

1 0 0 ω ω cos α 1 ω cos α 2<br />

⎜<br />

⎝ 0 1 0 ⎟<br />

⎠ = ⎜<br />

⎝ 0 ω sin α 1 ω cos α 3 sin α 2<br />

⎟<br />

⎠ (4.73)<br />

0 0 1 0 0 ω sin α 2 sin α 3<br />

The geometrical interpretation of this shape transformation can be readily read from the<br />

above example: the angle between the unit vectors ˆx, ŷ and ẑ after this transformation<br />

becomes<br />

cos (ˆx S , ŷ S ) = cos α 1<br />

cos (ˆx S , ẑ S ) = cos α 2<br />

cos (ŷ S , ẑ S ) = cos α 1 cos α 2 + sin α 1 sin α 2 cos α 3<br />

We can now consider a transformation like this X ⃗ RS = RSX ⃗ This is a big general class<br />

of models and for simplicity, we will look at two special families following Bowen &<br />

Ferreira (2002).<br />

• Family One: this is when we squeeze the cube in one direction and keep the other<br />

two angles orthogonal.<br />

α 1 = α ∈ [0, π/2]; α 2 = α 3 = π/2 (4.74)


4.3 A BIPOLAR POWER SPECTRUM STUDY OF COSMIC TOPOLOGY 61<br />

R 1 = R 2 = R 3 = R<br />

The transformation matrix then will be<br />

X RS =<br />

⎛<br />

⎞<br />

Rω Rω cos α 0<br />

⎜<br />

⎝ 0 Rω sin α 0 ⎟<br />

⎠ (4.75)<br />

0 0 Rω<br />

• Family Two, is the family of models in which all three angles are squeezed, but<br />

they are equal,<br />

α 1 = α 2 = α 3 = α ∈ [0, π/2]; (4.76)<br />

R 1 = R 2 = R 3 = R.<br />

The transformation matrix for this family is<br />

X S =<br />

⎛<br />

⎞<br />

Rω Rω cos α Rω cos α<br />

⎜<br />

⎝ 0 Rω sin α Rω cos α sin α ⎟<br />

⎠ (4.77)<br />

0 0 Rω sin α 2<br />

To calculate the analytical expression for BiPS for these models, we should repeat our<br />

previous steps. The correlation function for these models is the one for torus, but <strong>with</strong><br />

different wave modes,<br />

where k ⃗n<br />

= 2π L<br />

C T 3 (ˆq 1 , ˆq 2 ) =<br />

up <strong>with</strong> the squeezing angles.<br />

1 ∑<br />

P<br />

L 3 Φ (k ⃗n ) exp [−i ⃗ k ⃗n · (ˆq 1 − ˆq 2 )R SLS ] (4.78)<br />

⃗n<br />

⃗n. The wavenumbers are quantized and the effect of S is to mix them<br />

The ⃗ k vectors get transformed by the inverse of our<br />

transformation matrix, ⃗ k RS = S −1 R −1 ⃗ k, and hence is given by<br />

k 1 = 2π [<br />

n 1 − n 2 cot α 1 + n 3 (cot α 1 cot α 3 − cot α ]<br />

2<br />

) , (4.79)<br />

R 1 ω<br />

sin α 3<br />

k 2 = 2π [<br />

]<br />

1 cot α 3<br />

n 2 − n 3 ,<br />

R 2 ω sin α 1 sin α 1<br />

k 3 = 2π [<br />

]<br />

1<br />

n 3 .<br />

R 3 ω sin α 2 sin α 3


4.3 A BIPOLAR POWER SPECTRUM STUDY OF COSMIC TOPOLOGY 62<br />

Our aim here is to compute the BiPS analytically for the lowest wavenumbers, say for<br />

⃗n · ⃗n = 1.<br />

which corresponds to the following values of ⃗ k<br />

n 1 = ±1 , n 2 = n 3 = 0 =⇒ k 1 = ± 2π<br />

R 1 ω , k 2 = k 3 = 0, (4.80)<br />

n 2 = ±1 , n 1 = n 3 = 0 =⇒ k 1 = ± 2π<br />

R 1 ω cot α 1 , k 2 = ± 2π<br />

R 1 ω<br />

1<br />

sin α 1<br />

, k 3 = 0,<br />

n 3 = ±1 , n 1 = n 2 = 0 =⇒ k 1 = ± 2π<br />

R 1 ω (cot α 1 cot α 3 − cot α 2<br />

) , k 2 = ± 2π<br />

sin α 3 R 2 ω<br />

1<br />

k 3 = ± 2π<br />

R 3 ω<br />

sin α 2 sin α 3<br />

cot α 3<br />

sin α 1<br />

,<br />

The full calculation will be very tedious, and we will limit ourselves to Family One for<br />

simplicity<br />

The correlation function is then given as<br />

R 1 = R 2 = R 3 = R (4.81)<br />

α 2 = α 3 = π/2; α 1 = α ∈ [0, π/2]<br />

C (1) (ˆq 1 , ˆq 2 ) = C 0<br />

6<br />

= C 0<br />

3<br />

[ exp (−iɛ∆q 1 ) + exp (iɛ∆q 1 ) (4.82)<br />

+ exp −iɛ(cot α∆q 1 + 1<br />

sin α ∆q 2) + exp iɛ(cot α∆q 1 + 1<br />

sin α ∆q 2)<br />

+ exp −iɛ∆q 3 + exp iɛ∆q 3 ]<br />

[ cos ɛ∆q 1 + cos (ɛ cot α∆q 1 + α<br />

sin α ∆q 2) + cos ɛ∆q 3 ].<br />

where<br />

C 0 = 6 L 3 P Φ(k 1 ), (4.83)<br />

ɛ = 2π<br />

Rω 2R SLS sin θ/2<br />

And after expanding this expression in powers of ɛ, the leading order approximation to<br />

correlation C(ˆq, ˆq ′ ) turns out to have the form<br />

[<br />

3∑<br />

C (1) (ˆq 1 , ˆq 2 ) ≈ C 0 1 − ɛ2<br />

∆q 2 ɛ 2<br />

3 sin α 2 i +<br />

3 sin α cos α ∑ ]<br />

γ 2 ij ∆q i ∆q j . (4.84)<br />

i=1<br />

i≠j


4.3 A BIPOLAR POWER SPECTRUM STUDY OF COSMIC TOPOLOGY 63<br />

in which<br />

γ ij =<br />

⎛<br />

⎞<br />

0 1 0<br />

⎜<br />

⎝ 1 0 0 ⎟<br />

⎠ (4.85)<br />

0 0 cos α<br />

We can compute the BiPS for this correlation function analytically, but it is interesting<br />

to speculate what to expect from the analytical calculations.<br />

correlation can be written in this simple way<br />

C (1) (ˆq 1 , ˆq 2 ) ≈ C 0<br />

[<br />

1 − ɛ 2 |∆q|2 + cos 2 α∆q 2 3<br />

3 sin 2 α<br />

The above two point<br />

+ 2ɛ 2 cos α∆q 1 ∆q 2<br />

]<br />

. (4.86)<br />

The second term on the RHS of the above expression is our Toy Model I (<strong>with</strong> three<br />

equal axes), the third term is Toy Model I <strong>with</strong> one axes and the fourth term is the toy<br />

model <strong>with</strong> cross terms,<br />

C (3)<br />

T 3 = Toy Model I With Three Equal Axes ( O(ɛ 2 ) ) (4.87)<br />

+ Toy Model I With One Axes ( O(ɛ 2 ) )<br />

+ Cross Terms ( O(ɛ 2 ) )<br />

The first term is rotationally invariant and has no non-trivial component of BiPS, but<br />

second and third terms have a l = 2 components of BiPS which scales as O(ɛ 2 ). We<br />

can analytically calculate the BiPS for these models to be<br />

κ 0<br />

C0<br />

2<br />

κ 2<br />

C0<br />

2<br />

= π 2 [1 + 8 sin 2 α + 64<br />

27 cos4 α − 12ɛ 2 (1 − 16<br />

27 cos2 α)]<br />

= 64π2<br />

135 ɛ4 (cos 4 α + 3 cos2 α<br />

sin 4 ). (4.88)<br />

α<br />

BiPS has a non-zero l = 2 component as we expected. κ 2 vanishes for very large space<br />

(ɛ → 0).<br />

4.3.2 Spherical Spaces<br />

Flat spaces are not the only possibilities for the the topology of the universe. Recent<br />

observations suggest that the universe is “very close” to flat <strong>with</strong> a total density very<br />

close to one, Ω tot = 1.02 ± 0.02, Bennett et al. (2003); Spergel et al. (2003). This


4.3 A BIPOLAR POWER SPECTRUM STUDY OF COSMIC TOPOLOGY 64<br />

leaves the possibility of having “nearly flat” compact spaces as a candidate for the<br />

topology of our universe. This means besides the 18 possible Euclidean spaceforms (see<br />

e.g. Riazuelo et al. (2004b)) we have a whole set of countable infinity of spherical<br />

spaceforms (see Gausmann et al. (2001)) and a non-countable infinity of hyperbolic<br />

spaceforms (see Weeks (2003)). Weeks et al. (2004) showed that only a special<br />

family of closed multi-connected spaces called “well-proportioned” spaces, make the<br />

best candidates for a topological explanation of the low <strong>CMB</strong> quadrupole observed by<br />

COBE and WMAP (Fig. 4.11). Well-proportioned spaces are spaces whose dimensions<br />

are approximately equal in all directions. We studied some of these spaces in the previous<br />

section. Recently spherical spaces have attracted a lot of attention, see e.g., Luminet<br />

et al. (2003); Aurich et al. (2005); Roukema et al. (2004); Lachieze-Rey (2004);<br />

Luminet (2005). Motivated by these, we study multi-connected spaces <strong>with</strong> spherical<br />

geometry in the following sections. The universal covering space of these topologies is<br />

the 3-sphere, S 3 .<br />

Figure 4.11 <strong>CMB</strong> temperature correlation function of the WMAP and COBE data (left)<br />

and of the Poincare dodecahedron (right). Curves correspond to Poincare dodecahedron<br />

spaces <strong>with</strong> Ω tot = 1.0001 (red), 1.013 (green), 1.03 (blue) and 1.2 (purple). Figures<br />

from Bennett et al. (2003) and Hajian et al. (2006).


4.3 A BIPOLAR POWER SPECTRUM STUDY OF COSMIC TOPOLOGY 65<br />

Classification<br />

The isometry the 3-sphere is SO(4) and every isometry in SO(4) can be decomposed<br />

as the product of a right-handed and a left-handed Clifford translation. The 3-sphere<br />

S 3 enjoys a group structure as the set Q 3 of unit length quaternions. Each right(left)-<br />

handed Clifford translation corresponds to left (right) quaternion multiplication of Q 3 , so<br />

the group of right(left)-handed Clifford translations is isomorphic to Q 3 . It follows that<br />

SO(4) is isomorphic to Q 3 × Q 3 /{±(1, 1)} and thus the classification of the subgroups<br />

of SO(4) can be obtained from the classification of subgroups of Q 3 which are<br />

• the cyclic groups Z n of order n,<br />

• the binary 3 dihedral groups D ∗ m<br />

of order 4m, m ≥ 2,<br />

• the binary tetrahedral group T ∗ of order 24,<br />

• the binary octahedral group O ∗ of order 48,<br />

• the binary icosahedral group I ∗ of order 120 Riazuelo et al.<br />

(2004a),<br />

From the above classification, we can show that there are three categories of spherical<br />

three manifolds, these categories are briefly given below and in table 4.2. A detail<br />

description of spherical spaces is given in Gausmann et al. (2001).<br />

• The single action manifolds are those for which a subgroup R (resp. L) of Q 3<br />

acts as pure right-handed (resp. pure left-handed) Clifford translations. They<br />

are thus the simplest spherical manifolds and can all be written as S 3 /Γ <strong>with</strong><br />

Γ = Z n , Dm, ∗ T ∗ , O ∗ , I ∗ .<br />

• The double-action manifolds are those for which subgroups R and L of Q 3 act<br />

simultaneously as right- and left-handed Clifford translations, and every element<br />

of R occurs <strong>with</strong> every element of L. There are obtained for the groups Γ =<br />

Γ 1 × Γ 2 <strong>with</strong> (Γ 1 , Γ 2 ) = (Z m , Z n ), (Dm ∗ , Z n), (T ∗ , Z n ), (O ∗ , Z n ), (I ∗ , Z n ), <strong>with</strong><br />

gcd(m, n) = 1, gcd(4m, n) = 1, gcd(24, n) = 1, gcd(48, n) = 1, and gcd(120, n) =<br />

1, respectively.<br />

3 A binary group is the two-fold cover of the corresponding plain group.


4.3 A BIPOLAR POWER SPECTRUM STUDY OF COSMIC TOPOLOGY 66<br />

Name of the manifold Symbol Order Fundamental domain<br />

cyclic Z n n lens<br />

binary dihedral Dm ∗ 4m 2m-sided prism<br />

binary tetrahedral T ∗ 24 octahedron<br />

binary octahedral O ∗ 48 truncated cube<br />

binary icosahedral I ∗ 120 dodecahedron<br />

Table 4.2 Finite subgroups of Q 3 . There is a two-to-one homomorphism from Q 3 to<br />

SO(3) so the finite subgroups of SO(3) – the cyclic, dihedral, tetrahedral, octahedral<br />

and icosahedral groups – lift to Q 3 . Each such group lifts two-to-one, giving the corresponding<br />

binary group. Only the cyclic groups of odd order lift one-to-one. That is,<br />

each even-order cyclic group Z 2n ⊂ Q 3 arises as the two-fold cover of Z n ⊂ SO(3), while<br />

each odd-order cyclic group Z n ⊂ Q 3 is the one-to-one lift of Z n ⊂ SO(3). The binary<br />

polyhedral groups are not merely the product of the original polyhedral groups <strong>with</strong><br />

a Z 2 factor, nor are they isomorphic to the so-called extended polyhedral groups containing<br />

orientation-reversing as well as orientation-preserving isometries; rather they are<br />

something completely new. The plain dihedral, tetrahedral, octahedral and icosahedral<br />

groups do not occur as subgroups of Q 3 . From Weeks et al. (2003).<br />

• The linked-action manifolds are similar to the double action manifolds, except that<br />

each element of R occurs <strong>with</strong> only some of the elements of L.<br />

Single-action manifolds are the simplest class of spherical 3-manifolds and are directly<br />

given by finite subgroups of Q 3 . They are all given below<br />

• Each cyclic group Z n gives a lens space L(n, 1), whose fundamental domain is a<br />

lens shaped solid, n of which tile the 3–sphere.<br />

• Each binary dihedral group Dm ∗ gives a prism manifold, whose fundamental domain<br />

is a 2m–sided prism, 4m of which tile the 3–sphere.<br />

• The binary tetrahedral group T ∗ gives the octahedral space, whose fundamental<br />

domain is a regular octahedron, 24 of which tile the 3–sphere in the pattern of a<br />

regular 24–cell.


4.3 A BIPOLAR POWER SPECTRUM STUDY OF COSMIC TOPOLOGY 67<br />

• The binary octahedral group O ∗ gives the truncated cube space, whose fundamental<br />

domain is a truncated cube, 48 of which tile the 3–sphere.<br />

• The binary icosahedral group I ∗ gives the Poincaré dodecahedral space, whose<br />

fundamental domain is a regular dodecahedron, 120 of which tile the 3–sphere in<br />

the pattern of a regular 120–cell. Poincaré discovered this manifold in a purely<br />

topological context, as the first example of a multiply connected homology sphere.<br />

A quarter century later Weber and Seifert glued opposite faces of a dodecahedron<br />

and showed that the resulting manifold was homeomorphic to Poincaré’s homology<br />

sphere Gausmann et al. (2001).<br />

The other two categories are also classified as given in Gausmann et al. (2001). But we<br />

are only interested in Poincaré dodecahedral spaces and hence will stop right here. What<br />

we need to compute two point correlations in these spaces is the matrix representation<br />

of the icosahedral group, I, which generates the Poincaré dodecahedral space. This<br />

matrix is given in the appendix B of Gausmann et al. (2001) and we quote them here.<br />

Matrix representation of Icosahedral group, I<br />

The icosahedron’s twelve vertices being at (±φ, ±1, 0), (0, ±φ, ±1), and (±1, 0, ±φ),<br />

where φ = ( √ 5 − 1)/2 is the golden ratio, the icosahedral group consists of<br />

• the identity,<br />

• order 2 rotations about the midpoints of the icosahedron’s edges, realized as π rotations<br />

about the fifteen axes (1,0,0), (0,1,0), (0,0,1), (±φ, ±(φ+1), 1), (1, ±φ, ±(φ+<br />

1)), and (±(φ + 1), 1, ±φ), and<br />

• order 3 rotations about the centers of the icosahedron’s faces, realized as counterclockwise<br />

2π/3 rotations about the twenty axes (±1, ±1, ±1), (0, ±(φ + 2), ±1),<br />

(±1, 0, ±(φ + 2)), (±(φ + 2), ±1, 0), and<br />

• order 5 rotations about the icosahedron’s vertices, realized as both counterclockwise<br />

2π/5 rotations and counterclockwise 4π/5 rotations about the twelve axes<br />

(±φ, ±1, 0), (0, ±φ, ±1), and (±1, 0, ±φ).


4.3 A BIPOLAR POWER SPECTRUM STUDY OF COSMIC TOPOLOGY 68<br />

Figure 4.12 Fundamental domain for the regular dodecahedron which corresponds to<br />

the spaces generated by the binary icosahedral group, I. All face-pairings are Clifford<br />

translations, which implies that each face is glued to the face opposite, <strong>with</strong> a twist<br />

equal to the translation distance.<br />

This defines a total of 1 + 15 + 20 + 24 = 60 = |I| distinct rotations. These yield the<br />

120 quaternions of the binary icosahedral group I ∗ .<br />

BiPS of Poincaré dodecahedral spaces<br />

Simulating the <strong>CMB</strong> temperature anisotropy in Poincaré dodecahedral spaces requires<br />

knowing the eigenmodes of the Laplace operator on the space. These eigenmodes<br />

are given in Lachieze-Rey (2004); Weeks (2005). An alternative method developed<br />

by Hipolito-Ricaldi & Gomero (2005) expresses the covariance matrix in terms of the<br />

elements of the covering group of the space, Γ, instead of computing the eigenmodes<br />

of the Laplacian operator. However, the universal cover, S 3 , of the Poincaré dodecahedral<br />

spaces is a very simple manifold whose eigenmodes are very well studied and two<br />

point correlations on this space is very easily computable. The regularized method of<br />

images developed by Bond, Pogosyan & Souradeep (1998, 2000), allows computation<br />

of the correlation function ξ [Γ] (x, x ′ ) on a Poincaré dodecahedral space where Γ is the


4.3 A BIPOLAR POWER SPECTRUM STUDY OF COSMIC TOPOLOGY 69<br />

Figure 4.13 Two point correlations in Poincaré dodecahedral spaces, C(North P ole, ˆn).<br />

Correlations have been computed numerically using the method of images including<br />

only naive Sachs-Wolfe effect for Ω k = 0.0001 (top left), 0.013 (top right), 0.2 (bottom<br />

left). Hot spots represent the highly correlated regions. Inclusion of the integrated<br />

Sachs-Wolfe effect weakens the correlations, the right panel on bottom shows the two<br />

point correlation computed <strong>with</strong> nSW+ISW for a space <strong>with</strong> Ω k = 0.013.<br />

Icosahedral group discussed in the previous section. Right panel of Fig. 4.11 shows the<br />

two point correlations of Poincaré dodecahedral spaces computed for four different sizes<br />

of the space. The power gets suppressed on large angular scales as we go to smaller<br />

spaces. This feature of Poincaré dodecahedral spaces has been used to explain the low<br />

quadrupole in the WMAP data by some authors, (see e.g. Luminet et al. (2003)). But<br />

this suppression of the quadrupole happens at the cost of introducing highly correlated<br />

spots which are spatially far apart in the <strong>CMB</strong> sky. Some examples of this have been<br />

presented in Fig. 4.13, in which we have plotted the correlations for a fixed point of<br />

the sky (North Galactic Pole) and the rest of the sky in Poincaré dodecahedral spaces.<br />

When the space is very big (when it is very close to flat, e.g. Ω tot = 1.0001), the correlation<br />

falls off rapidly as we go further from the Pole. But in smaller spaces such as<br />

the one corresponding to Ω tot = 1.013, regions <strong>with</strong> strong correlations appear on the


4.3 A BIPOLAR POWER SPECTRUM STUDY OF COSMIC TOPOLOGY 70<br />

map. These spots are signatures of small spaces and as we discussed in the section on<br />

flat spaces, they are signs of violation of statistical isotropy in compact spaces. These<br />

correlations are suppressed as we add the integrated Sachs-Wolfe effect to the computation<br />

of two point correlations, but they do not disappear (right panel on the bottom of<br />

Fig. 4.13). Using the fact that statistical isotropy is violated in Poincaré dodecahedral<br />

spaces one could use the bipolar power spectrum as a probe to look for the topology of<br />

the universe. We compute the BiPS from the correlation functions in real space using<br />

eqn. (2.30). Fig. 4.14 plots the BiPS for a number of Poincaré dodecahedral spaces. We<br />

find the following interesting results:<br />

(a)<br />

(b)<br />

(c)<br />

κ l = 0 for odd l for all Poincaré dodecahedral spaces. This feature has been<br />

studied by Mukhopadhyay (2005) and can be explained by 2n-fold rotational<br />

symmetries of the space about an axis.<br />

The first non-zero BiPS component is the l = 6 component. This is a clear<br />

signature of Poincaré dodecahedral spaces.<br />

Other components of BiPS become non-zero as the inner radius, R < , decreases<br />

and BiPS signature shifts to larger l for small spaces.<br />

Above we presented the theoretical predictions of BiPS signatures of Poincaré dodecahedral<br />

spaces. To investigate the detectability of non-zero BiPS from the data, we<br />

simulated 1000 <strong>CMB</strong> temperature anisotropy maps for various Poincaré dodecahedral<br />

spaces at HEALPix resolution of N side = 16 corresponding to l max = 47. Methods<br />

which use the angular power spectrum to simulate <strong>CMB</strong> maps (like HEALPix) can not<br />

be used in simulating <strong>CMB</strong> maps for compact spaces. This can instead be done using<br />

the methods presented in Appendix F to simulate the <strong>CMB</strong> anisotropy maps from theoretical<br />

two point correlation matrices. We added noise maps to these <strong>CMB</strong> simulated<br />

maps. Noise maps were simulated from the coadded noise covariance matrix of WMAP.<br />

Since BiPS is orientation independent, we don’t need to worry about the orientation of<br />

the Dirichlet domain <strong>with</strong> respect to the <strong>CMB</strong> map. We computed the BiPS for these<br />

simulated <strong>CMB</strong> anisotropy maps for various Poincaré dodecahedral spaces using the<br />

estimator given in eqn. (2.37). We average these 1000 BiPS and compute the dispersion<br />

in them. The dispersion is an estimate of 1σ error bars. At last we use the theoretical


4.3 A BIPOLAR POWER SPECTRUM STUDY OF COSMIC TOPOLOGY 71<br />

angular power spectrum of simulated maps to estimate the bias and correct for it. We<br />

have plotted the BiPS of three models in Fig. 4.16. Which shows what we should expect<br />

to get from the data if the observed universe was a Poincaré dodecahedral space. This<br />

is used by Hajian et al. (2006) to put constraints on Poincaré dodecahedral topologies.


4.3 A BIPOLAR POWER SPECTRUM STUDY OF COSMIC TOPOLOGY 72<br />

3.5 x 1010 l<br />

7 x 1010 l<br />

3<br />

6<br />

2.5<br />

5<br />

2<br />

4<br />

κ l<br />

κ l<br />

1.5<br />

3<br />

1<br />

2<br />

0.5<br />

1<br />

0<br />

0 2 4 6 8 10 12 14 16 18 20<br />

0<br />

0 2 4 6 8 10 12 14 16 18 20<br />

6 x 1010 l<br />

3 x 1010 l<br />

5<br />

2.5<br />

4<br />

2<br />

κ l<br />

3<br />

κ l<br />

1.5<br />

2<br />

1<br />

1<br />

0.5<br />

0<br />

0 2 4 6 8 10 12 14 16 18 20<br />

0<br />

0 2 4 6 8 10 12 14 16 18 20<br />

Figure 4.14 BiPS of Poincaré dodecahedron spaces for Ω k = 0.013 (top left), Ω k = 0.03<br />

(top right), Ω k = 0.1 (bottom left) and Ω k = 0.2 (bottom right). Computed from a<br />

theoretical two point correlation at HEALPix resolution N side = 16. Figures from Hajian<br />

et al. (2006).


4.3 A BIPOLAR POWER SPECTRUM STUDY OF COSMIC TOPOLOGY 73<br />

Figure 4.15 Simulated <strong>CMB</strong> maps of Poincaré dodecahedral spaces for Ω k =<br />

0.013, 0.03 and 0.1.


4.3 A BIPOLAR POWER SPECTRUM STUDY OF COSMIC TOPOLOGY 74<br />

0<br />

5 10 15<br />

0<br />

5 10 15<br />

0<br />

5 10 15<br />

Figure 4.16 Average of BiPS of simulated <strong>CMB</strong> maps of Poincaré dodecahedral spaces<br />

obtained from 1000 simulations for Ω k = 0.013, 0.03 and 0.1, Smoothed <strong>with</strong> W G (l s =<br />

18). Simulated noise maps from coadded noise covariance matrix of WMAP has been<br />

added to the <strong>CMB</strong> maps. Figures from Hajian et al. (2006).


Chapter 5<br />

Homogeneous Cosmological Models<br />

Friedmann-Robertson-Walker (FRW) models are simplest models of the expanding<br />

Universe which are spatially homogeneous and isotropic. Although the principal<br />

justification for studying these models was mathematical tractability, rather than observational<br />

evidence. But now we have strong observational evidence from isotropy of<br />

<strong>CMB</strong> and large scale structure of the universe that the structure of the universe at<br />

large scales must be very close to that of a FRW model. But there is still a freedom to<br />

choose homogeneous models which are initially anisotropic and as the time goes on they<br />

become more isotropic and will asymptotically tend to FRW models. Bianchi models<br />

provide a generic description of homogeneous anisotropic cosmologies and they are classified<br />

into 10 equivalence classes Ellis & MacCallum (1969). Among these models it is<br />

reasonable to consider only those Bianchi types which admit FRW models. These are<br />

types I and VII 0 in the case of k = 0, V and VII h in the case of k = −1, and IX in<br />

the case of k = +1. Bianchi models which do not admit FRW solutions become highly<br />

anisotropic at large times. Type IX models recollapse after a finite time and hence will<br />

not approach arbitrarily near to isotropy. Also models of type VII h will not in general<br />

approach isotropy Collins & Hawking (1973b). The most general Bianchi types which<br />

admit FRW models are Bianchi types VII h and IX. These two types contain types I, V,<br />

VII 0 as special subcases. An interesting feature of these models is that they resemble<br />

a universe <strong>with</strong> a vorticity. It is interesting to determine bounds on universal rotation<br />

from cosmological observations because the absence of such a rotation is a prediction of<br />

inflation.<br />

75


5.0 76<br />

<strong>CMB</strong> anisotropy is a strong tool to study the evolution of vorticity in the universe<br />

because such models have clear <strong>CMB</strong> profiles. These patterns which are imprinted<br />

by the unperturbed anisotropic expansion are roughly constructed out of two parts:<br />

production of pure quadrupole variations (or focusing of the quadrupole pattern into a<br />

hotspot in open models) and a spiral pattern which is the signature of VII 0 and VII h<br />

models Collins & Hawking (1973b); Doroshkevich et al. (1975); Barrow et al. (1985).<br />

This temperature anisotropy pattern for Bianchi VII h universes is of the form<br />

∆T B<br />

(θ, φ) = f 1 (θ) sin φ + f 2 (θ) cos φ. (5.1)<br />

T 0<br />

The functions f 1 (θ) and f 2 (θ) depend on details of the particular Bianchi model and<br />

must be computed numerically Barrow et al. (1985). There are distinct features in<br />

each Bianchi model. In a pioneering work analytical arguments were used to find upper<br />

bounds on the amount of shear and vorticity in the universe today, from the absence<br />

of any detected <strong>CMB</strong> anisotropy Collins & Hawking (1973b). A detailed numerical<br />

analysis of such models used experimental limits on the dipole and quadrupole to refine<br />

limits on universal rotation Barrow et al. (1985). After the first detection of <strong>CMB</strong><br />

anisotropy by COBE-DMR, Bianchi models were again studied by fitting the full spiral<br />

pattern from models <strong>with</strong> global rotation to the 4-year DMR data to constrain the<br />

allowed parameters of a Bianchi model of type VII h Bunn et al. (1996); Kogut et al.<br />

(1997). Recently Bianchi VII h models were compared to the first-year WMAP data on<br />

large scales and it was shown that the best fit Bianchi model corresponds to a highly<br />

hyperbolic model (Ω 0 = 0.5) <strong>with</strong> a right-handed vorticity ( ω ) H 0 = 4.3 × 10 −10 (Jaffe et<br />

al. (2005)). These models may be ruled out by their low density but it is curious to<br />

note that Bianchi models are examples of breakdown of statistical isotropy even <strong>with</strong>out<br />

considering a perturbed universe. Hence tests of SI can be carried out to put constraints<br />

on various parameters of these models Ferreira & Magueijo (1997). As an example,<br />

let’s assume that a <strong>CMB</strong> map is constructed out of two ingredients<br />

∆T (ˆn) = ∆T I (ˆn) + ∆T B (ˆn), (5.2)<br />

where ∆T B is the Bianchi pattern and ∆T I represents the isotropic residual fluctuation<br />

caused by variations in the density and gravitational potential of a FRW universe.


5.1 CLASSIFICATION OF COSMOLOGICAL MODELS 77<br />

The BiPS due to the Bianchi template can be computed for different models. As it<br />

is shown in fig. 5.1, different Bianchi models have different characteristic BiPS that<br />

can be used to constrain them using the first-year WMAP data. In summary, if the<br />

<strong>CMB</strong> sky is a superposition of a SI part and a Bianchi pattern, the resultant <strong>CMB</strong><br />

map should be statistically anisotropic and in principle, that should be seen by BiPS.<br />

Exact bounds on these models from BiPS based on detailed studies, are given at the<br />

final section of this chapter. Also as it was discovered by Jaffe et al. (2005), the<br />

deviations from Gaussianity in the kurtosis of spherical Mexican hat wavelet coefficients<br />

are eliminated once the data is corrected for the Bianchi template. In a recent work<br />

McEwen et al. (2005) investigated the effect of this Bianchi correction on the detections<br />

of non-Gaussianity in the WMAP data. According to McEwen et al. (2005) previous<br />

detections of non-Gaussianity observed in the skewness of spherical wavelet coefficients<br />

reported in McEwen et al. (2005a) are not reduced by the Bianchi correction and<br />

remain at a significant level of 98.4%.<br />

In the next section (and in Appendix D) we will study the symmetries of the space<br />

to show how spatially homogeneous models are constructed. Most of definitions and<br />

mathematical notations in the following sections are borrowed from Nakahara (1984),<br />

Wald (1984) and Ellis & van Elst (1999).<br />

5.1 Classification of Cosmological Models<br />

We can classify cosmological models by their symmetries. For any space, the dimension<br />

of the group of symmetries of the space (group of isometries), is given by the sum<br />

of dimensions of groups of rotational and translational symmetries 1 , i.e.,<br />

dim group of isometries (r) = dim group of rotational symmetries (s) + dim group of<br />

translational symmetries (q).<br />

In a 4-dimensional cosmological model, r can pick up different values. These values<br />

can be obtained by a variety of q and s. The possibilities for the dimension of orbits, s,<br />

are 0, 1, 2, 3 and 4. For isotropies of the spatial dimensions, q can be 0, 1 or 3. q = 2 is<br />

excluded because in cosmology we consider non-empty perfect fluids in which ρ + p > 0<br />

1 For a detailed mathematical description of this see Appendix D.


5.1 CLASSIFICATION OF COSMOLOGICAL MODELS 78<br />

Figure 5.1 BiPS of Bianchi VII h models (arbitrary units); for a hyperbolic model x =<br />

0.55, Ω 0 = 0.5 (left) and a nearly flat model x = 0.55, Ω 0 = 0.99 (right). If the observed<br />

<strong>CMB</strong> map contains a pattern in it, it will violate the statistical isotropy which would<br />

be seen in the BiPS.<br />

and hence there will be uniquely defined notions of the average velocity of matter, and<br />

corresponding preferred world-lines. Their 4-velocity in each case is given by<br />

u µ = dx µ /dτ, (5.3)<br />

where τ is proper time measured along the fundamental world-lines. We assume this<br />

4-velocity is unique; that is, there is a well-defined preferred motion of matter at each<br />

space-time event. This u µ is therefore invariant and hence allowed rotations are those<br />

which act orthogonally on u µ . There is no 2-dimensional subgroup of O(3) and hence<br />

q = 2 is excluded from the allowed values that q can choose. All over the space, r should<br />

be fixed. But q can vary. For example right at the axis center of symmetry, q can be<br />

bigger. But at that point the dimension of orbits is less and hence r remains fixed.<br />

Given these, we can have the following spaces:<br />

(i)<br />

Possibilities for isotropy;<br />

1. Isotropic: q = 3,<br />

2. Local Rotational Symmetry: q = 1,<br />

3. Anisotropic: q = 0;


5.2 CLASSIFICATION OF COSMOLOGICAL MODELS 79<br />

Table 5.1.<br />

Classification of Cosmological Models <strong>with</strong> Respect to Their Symmetries.<br />

... inhomogeneous spatially homogeneous space-time homogeneous<br />

... (s=2) (s=3) (s=4)<br />

Anisotropic spatially self-similar, Bianchi Osvath/Kerr<br />

(q = 0) Admitting Abelian<br />

or non-Abelian G 2<br />

Local Rotational Symmetry Lemaitre-Tolman- Kantowski-Sachs Godel stationary<br />

(q = 1) Bondi family rotating universe<br />

Isotropic not allowed* FLRW Einstein static<br />

(q = 3) universe<br />

Note. — Table reproduced from Ellis & van Elst (1999).<br />

∗ inhomogeneous models cannot be isotropic about a general point because isotropy everywhere implies<br />

spatial homogeneity;<br />

(ii)<br />

Possibilities for homogeneity;<br />

1. Space-time homogeneous models: s = 4<br />

2. Spatially homogeneous universes: s = 3<br />

3. Spatially inhomogeneous universes: s ≤ 2.<br />

Using the above classes, one can make every cosmological model <strong>with</strong> a given symmetry.<br />

Table 5.2 shows different varieties of cosmological models that may be constructed from<br />

different combinations of symmetries of the space. A complete list <strong>with</strong> more discussions<br />

is given in Ellis & van Elst (1999).<br />

For example the family of FLRW spaces that model the standard cosmology, are<br />

isotropic and spatially homogeneous universes (q = 3, s = 3 =⇒ r = 6). Another<br />

interesting family is the spatially homogeneous but anisotropic case (q = 0, s = 3 =⇒<br />

r = 3) which is the family of Bianchi universes. This family has a simply transitive<br />

group of isometries G 3 and hence no continuous isotropy group. In the rest of this<br />

chapter we study these models in detail.


5.2 MODERN CLASSIFICATION OF BIANCHI MODELS 80<br />

5.2 Modern Classification of Bianchi Models<br />

Today, the Bianchi classification of 3-dimensional Lie algebras is no longer presented<br />

by the original Bianchi method (see Appendix E). Instead, a simpler scheme for classifying<br />

the equivalence classes of 3-dimensional Lie algebras is used. This scheme uses the<br />

irreducible parts of the structure constant tensor under linear transformations rather<br />

than the more complicated derived group approach of Lie and Bianchi. Following Ellis<br />

and MacCallum Ellis & MacCallum (1969), we decompose the spatial commutation<br />

structure constants Cbc a (eqn. D.9) into a tensor, nab , and a vector, a b<br />

Cbc a = ɛ dbc n ad + δc a a b − δb a a c (5.4)<br />

where ɛ abc is the 3-dimensional antisymmetric tensor and n ab and a b are defined as<br />

a b = 1 2 Ca ba (5.5)<br />

The structure constants C a bc<br />

n ab = 1 2 C(a cd ɛb)cd<br />

expressed in this way, clearly satisfy the first Jacobi identity,<br />

eqn. (D.10). Second Jacobi identity, eqn. (D.11) shows that a b must have zero<br />

contraction <strong>with</strong> the symmetric 2-tensor n ab ;<br />

C a e[b Ce cd] = 0 =⇒ nab a b = 0 (5.6)<br />

We can choose convenient basis to diagonalize n ab to attain n ab = diag(n 1 , n 2 , n 3 ) and<br />

to set a b = (a, 0, 0). The Jacobi identities are then simply equivalent to n 1 a = 0.<br />

Consequently we can define two major classes of structure constants<br />

Class A : a = 0,<br />

Class B : a ≠ 0.<br />

One can then classify further by the sign of the eigenvalues of n ab (signs of n 1 , n 2 and<br />

n 3 ). This classification is given in Table 5.2. Parameter h in class B, where a is non-zero,<br />

is defined by the scalar constant of proportionality in the following relation<br />

a b a c = h 2 ɛ bikɛ cjl n ij n kl . (5.7)<br />

In the case of diagonal n ab , the h factor has a simple form h = a 2 /(n 2 n 3 ).<br />

In addition to these, we can have two different Bianchi models


5.3 THE FIELD EQUATIONS 81<br />

Table 5.2.<br />

Classification of Homogeneous Cosmological Models into Ten Equivalence<br />

Classes<br />

Group Class Group Type Eigenvalues of n ab a Dimension FLRW Type ...<br />

I 0, 0, 0 0 0 k=0 Abelian<br />

II +, 0, 0 0 3 —<br />

VI 0 0, +, − 0 5 —<br />

A VII 0 0, +, + 0 5 k=0<br />

VIII −, +, + 0 6 —<br />

IX +, +, + 0 6 k=+1<br />

V 0, 0, 0 + 3 k=-1<br />

IV 0, 0, + + 5 —<br />

B VI h 0, +, − + 6 — h < 0<br />

III 0, +, − n 2 n 3 6 VII h <strong>with</strong> h = −1<br />

VII h 0, +, + + 6 k=-1 h > 0<br />

Note. — Table from Ellis & van Elst (1999) & Collins & Hawking (1973a).<br />

Orthogonal models, <strong>with</strong> the fluid flow lines orthogonal to the surfaces of homogeneity;<br />

which means the fluid 4-velocity U µ is parallel to the normal vectors n µ . In this case,<br />

the matter variables will be just the fluid density and pressure.<br />

Tilted models, <strong>with</strong> the fluid flow lines not orthogonal to the surfaces of homogeneity;<br />

i.e. the fluid 4-velocity is not parallel to the normals. Hence we also need the peculiar<br />

velocity of the fluid relative to the normal vectors. These components enter as further<br />

variables.<br />

Rotating models must be tilted and have a more complex behavior Ellis & van Elst<br />

(1999).<br />

5.3 The Field Equations<br />

We follow Collins & Hawking (1973a) and Collins & Hawking (1973b) to obtain<br />

the Einstein field equations for Bianchi models.<br />

We saw that the spatially homogeneous models admit a group of symmetries, G,


E A µ satisfy the relation E A µ;ν − EA ν;µ = CA BC EB µ EC ν , (5.10)<br />

5.3 THE FIELD EQUATIONS 82<br />

which is transitive on a family of 3-dimensional space-like hypersurfaces, S(t). We label<br />

these spaces by a time parameter t which is chosen to measure proper time along the<br />

geodesics normal to S(t). The normal to the hypersurface is given by n α = t ;α such that<br />

g αβ n α n β = −1 and ; means covariant derivative 2 . We define three invariant vector fields<br />

E µ A and its dual EA µ<br />

at each point p ∈ S and covectors E A µ<br />

basis satisfy<br />

We drag E µ A and EA µ<br />

the Lie derivative of the fields<br />

on the surfaces of homogeneity by dragging the tangent vectors E µ A<br />

(one-forms) under the group. The above dual<br />

E A µ Eµ B = δA B (5.8)<br />

and E A µ Eν A = δν µ + n µn ν .<br />

by the unit vector field n ν orthogonal to S(t), this is done by<br />

Now at every point p ∈ S(t), vector fields E µ A and EA µ<br />

L nν Eµ A = L n ν<br />

E µ A<br />

= 0. (5.9)<br />

are defined and are invariant<br />

under the 3-dimensional Lie group G (group of isometries of the space). Covector fields<br />

where C A BC<br />

are constants given by the structure constants of the Lie group G<br />

C A BC = Cµ νλ EA µ Eµ B Eν C (5.11)<br />

The metric of this model can be written in the following form<br />

g µν = −n µ n ν + g AB E A µ E B ν . (5.12)<br />

We split the matrix g AB<br />

(1968),<br />

into two parts: its volume and the distortion part,Misner<br />

g AB = e 2α ( e 2β) AB . (5.13)<br />

where the scalar α represents the volumetric expansion whilst β is symmetric, trace free<br />

3 × 3 matrix and hence e 2β is given by the series<br />

∞∑<br />

e 2β 1<br />

=<br />

r! (2β)r<br />

r=0<br />

2 Greek indices run from 0 to 3, Latin indices from 1 to 3. Latin indices may be lowered and raised<br />

by the 3 × 3 matrix g AB = g µν E µ A Eν B and its inverse gAB = g µν E A µ E B ν which depend only on t.


5.3 THE FIELD EQUATIONS 83<br />

. Now we define an orthonormal basis X ν µ<br />

such that<br />

X 0 µ = −n µ, X i µ = eα (e 2β ) iA E A µ . (5.14)<br />

where i, j, k, · · · = 1, 2, 3. The 0-th component of these basis changes sign on raising<br />

or lowering, but rest of the components remain unchanged. Ricci rotation coefficients<br />

Γ αβγ are obtained from the connection coefficients through δ αδ Γ δ βγ . For this basis these<br />

coefficients satisfy Γ δɛγ + Γ γɛδ and are given by<br />

Using these properties we will obtain<br />

Γ = 1 2<br />

X γ µ;ν = Γγ δɛ Xδ µ Xɛ ν (5.15)<br />

[<br />

Xγ[µ;ν] X µ δ Xν ɛ + X ɛ[µ;ν]X µ γ Xν δ − X δ[µ;ν]X µ ɛ Xν γ<br />

]<br />

. (5.16)<br />

This <strong>with</strong> eqn. (5.10) will determine the components of Ricci rotation coefficients<br />

Γ ij0 = −Γ 0ij = (∂ 0 α)δ ij + σ ij , (5.17)<br />

Γ i00 = −Γ 00i = 0,<br />

Γ i0j = −ν ij<br />

Γ ijk = 1 2 e−α (e β ) F A (e −β ) GB (e −β ) HC C A BC [δ F iδ Gj δ Hk + δ F k δ Gi δ Hj − δ F j δ Gk δ Hi ] .<br />

In the above expressions, ∂ 0 denotes differentiation by t and σ ij represents the shear of<br />

the normals n µ and is given by anticommutator of (∂ 0 e β ) and e −β<br />

σ ij ≡ 1 2 {(∂ 0e β ), e −β } ij (5.18)<br />

= 1 2<br />

[<br />

(∂0 e β )e −β + e −β (∂ 0 e β ) ] ij .<br />

The extent of commutativity of ∂ 0 e β and e −β is given by their commutator denoted by<br />

ν ij<br />

ν ij ≡ 1 2<br />

[<br />

(∂0 e β ), e −β] ij<br />

(5.19)<br />

Ricci tensor of the space Sin the induced metric is<br />

Rij ∗ = − 1 4 e−2α { 2 CBCC A DA(e C −β ) Bi (e −β ) Dj (5.20)<br />

+ CBCC A EF D (e −2β ) BE [2(e 2β ) DA (e −β ) Ci (e −β ) F j − (e −2β ) CF (e β ) Ai (e β ) Dj ]<br />

+ CABC A DE(e C −2β ) BE [(e β ) Ci (e −β ) Dj + (e β ) Cj (e −β ) Di ]}.


5.4 APPROACHING ISOTROPY 84<br />

We can now write the field equations for this basis Collins & Hawking<br />

filed equations are<br />

(1973a). The<br />

3(∂ 0 α) 2 − 1 2 σ ijσ ij + 1 2 R∗ = 8πT 00 , (5.21)<br />

e −α [(e −β σe β ) BA CBC A (e−β ) Ci − σ ij (e −β ) jC CAC A ] = 8πT 0i , (5.22)<br />

∂ 0 σ ij + 3(∂ 0 α)σ ij + [σ, ν] ij + Rij ∗ − 1 3 R∗ δ ij = 8π(T ij − 1 3 T kkδ ij ) (5.23)<br />

−6∂0 2 α − 9(∂ 0α) 2 − 3 2 σ ijσ ij − 1 2 R∗ = 8πT kk . (5.24)<br />

5.4 Approaching Isotropy<br />

We say a model approaches isotropy if at arbitrary large time scales it has the<br />

following properties<br />

(a)<br />

The universe should continue to expand forever, which means<br />

t → +∞ =⇒ α → +∞<br />

Also we need<br />

(b)<br />

(c)<br />

(d)<br />

Second condition is the isotropy condition. T 0i /T 00 represents an average<br />

velocity of the matter relative to the surface of homogeneity. The universe<br />

can only reach isotropy only if at large times this velocity vanishes<br />

t → +∞ =⇒ T 00 > 0<br />

and<br />

T 0i<br />

T 00 → 0<br />

<strong>Anisotropy</strong> in the locally measured Hubble parameter should tend to zero.<br />

If we define σ 2 = 1 2 σ ijσ ij , this condition will then translate into<br />

t → +∞ =⇒<br />

σ<br />

∂ 0 α → 0<br />

If we notice that the anisotropy observed in <strong>CMB</strong> due to a specific Bianchi<br />

model is somehow related to the change of β between the time of last scattering<br />

and the time it is observed, we will conclude that the amazing degree<br />

of isototropy of the <strong>CMB</strong>, will tell us that β should be very small,<br />

t → +∞ =⇒ β → β 0 where β 0 = const.


5.5 ENERGY MOMENTUM TENSOR AND VORTICITY VECTOR 85<br />

The following theorem which was proved by Collins & Hawking (1973a) shows that<br />

only 4 Bianchi types approach isotropy.<br />

Theorem 5.4.1 If the dominant-energy condition 3 and the positive-pressure criterion 4<br />

are satisfied, the universe can approach isotropy only if it is one of the types I, V , V II 0<br />

and V II 5 h .<br />

Since we know that our universe is isotropic to a very high extent, from now on we will<br />

only be interested in these models.<br />

5.5 Energy Momentum Tensor and Vorticity Vector<br />

In a matter dominated universe the dominant contribution to the energy-momentum<br />

tensor comes from non-relativistic matter,<br />

T µν = ρ m U µ U ν , (5.25)<br />

where ρ m is the matter density and U µ is the flow-vector. The conservation equation<br />

tells us that flow-lines satisfy the geodesic equation, hence<br />

∂ t (ρ m U 0 e 3α ) = −ρ m e α C A ABU C (e −2β ) BC . (5.26)<br />

in models of class A, right hand side of the above expression vanishes and will result in<br />

Class A : ρ m = a m (U 0 e 3α ) −1 (5.27)<br />

where a m is a constant. Using this constant, T 00 can be written as a m e −3α (1 + E m )<br />

where a m is constant in class A models and hence the kinetic energy associated <strong>with</strong> the<br />

peculiar velocity of the matter relative to the surfaces of homogeneity will be<br />

E m = U 0 − 1 = [ U A U B e −2α (e −2β ) AB + 1 ] 1/2<br />

− 1. (5.28)<br />

And the anisotropic part of the spatial components of the energy-momentum tensor will<br />

be<br />

T ij − 1 3 T kkδ ij = −a m e −3α ∂E m<br />

∂β ij<br />

. (5.29)<br />

3 T 00 ≥ |T αβ |<br />

4 T kk ≥ 0<br />

5 Spatially homogeneous cosmological models <strong>with</strong> cosmological constant were investigated by Wald<br />

(1983) and Goliath & Ellis (1998)


5.6 ENERGY MOMENTUM TENSOR AND VORTICITY VECTOR 86<br />

In a radiation dominated universe, the energy momentum tensor would be<br />

and the energy density of radiation obeys<br />

T µν = 4 3 ρ γU µ U ν + 1 3 ρ γg µν , (5.30)<br />

∂ t<br />

[<br />

ργ (U 0 ) 4/3 e 4α] = − 4 3 ρ γ(U 0 ) 1/3 C A ABU C e 2α (e −2β ) B C. (5.31)<br />

Right hand side will vanish in class A models and as a result<br />

Class A : ρ γ (U 0 ) 4/3 e 4α = a γ (5.32)<br />

Flow lines feel an acceleration caused by the radiation pressure<br />

4<br />

3 ρ γU µ<br />

;ν U ν = − 1 3 ρ γ;νh µν , where h µν ≡ g µν + U µ U ν . (5.33)<br />

Because of homogeneity ρ γ;ν = −∂ t ρ γ n ν . Therefore<br />

∂ t (ρ 1/4<br />

γ U A ) = ρ 1/4<br />

γ (U 0 ) −1 C B CA U BU D e −2α (e −2β ) C D. (5.34)<br />

The kinetic energy and the anisotropic part of the spatial components of the energymomentum<br />

tensor are then given by Hawking (1968);<br />

E γ = (U 0 ) 2/3 − 1 = [ U A U B e −2α (e −2β ) AB + 1 ] 1/3<br />

− 1 (5.35)<br />

T ij − 1 3 T kkδ ij = −2a γ e −4α ∂E γ<br />

∂β ij<br />

.<br />

In these models we are mainly interested in the vorticity<br />

ω = (g AB ω A ω B ) 1/2 = (ω i ω i ) 1/2 (5.36)<br />

The vorticity vector of the matter is defined as ω µ = 1 2 ηµνλρ U ν U λ;rho . Thus<br />

ω A = 1 2 e−3α ɛ ABC<br />

[ 1<br />

2 CD BCU D U 0 + U B ∂ 0 U C<br />

]<br />

, (5.37)<br />

ω 0 = 1 2 e−3α ɛ ABC C D BC U AU D<br />

For small, non-relativistic velocities, U 0 ∼ 1 and U A is small, so<br />

ω A ≃ 1 4 e−3α ɛ ABC C D BC U D. (5.38)


5.6 NULL GEODESICS 87<br />

5.6 Null Geodesics<br />

In general, all of our observations take place on our past light cone, which will<br />

develop many cusps and caustics at early times because of gravitational lensing, but is<br />

still locally generated by geodesic null rays. The information we receive comes to us<br />

along these null rays, <strong>with</strong> tangent vector<br />

K µ K µ = 0 (5.39)<br />

where K µ is the tangent vector to the null geodesic from the observer in a given direction.<br />

In the orthonormal basis Xµ α which was defined in the previous section, the null geodesic<br />

condition will become<br />

K 0 K 0 + K A K A = 0 (5.40)<br />

As we saw, the 0-th component of these basis changes sign on raising or lowering. So,<br />

K 0 = −K 0 =⇒ K 0 = (K A K A ) 1/2 (5.41)<br />

and<br />

K A = g AB K B = e 2α (e 2β ) AB K B (5.42)<br />

<strong>with</strong> K A = (K cos θ, K sin θ cos φ, K sin θ sin φ) where K is constant. For small anisotropies<br />

Barrow et al. (1985), i.e. β ij ≪ 1,<br />

g AB ≃ e 2α δ AB =⇒ K 0 ≃ Ke −α (5.43)<br />

The null geodesics satisfy the geodesic equation<br />

In the orthonormal basis, the geodesic equations are<br />

K A ≃ e 2α δ AB K B<br />

K µ;ν K ν = 0, (5.44)<br />

∂ t K A = CCA<br />

B K B K C<br />

. (5.45)<br />

K 0<br />

Using the small anisotropy condition, eqn. (5.43), the geodesic equation to fist order<br />

will be<br />

∂ t K A = 1 K CB CA K BK C e −α . (5.46)


5.6 NULL GEODESICS 88<br />

We introduce a new time parameter by dτ = e −α dt and we denote the differentiation<br />

<strong>with</strong> respect to this time by prime<br />

K A ′ = 1 K CB CAK B K C . (5.47)<br />

We are interested in those Bianchi types which are containing FLRW models as special<br />

isotropic cases. These models are Bianchi types I, V , V II 0 , V II h and XI. Null geodesic<br />

equations and their solutions for these models were first derived by Barrow et al. (1985)<br />

and are given below<br />

Bianchi Type I<br />

All structure constants are zero,<br />

θ ′ = φ ′ = 0 (5.48)<br />

=⇒ θ = θ 0 = const.<br />

φ = φ 0 = const.<br />

Bianchi Type V<br />

Non-zero structure constants:<br />

C 2 21 = −C 2 12 = C 3 31 = −C 3 13 = 1.<br />

θ ′ = − sin θ, φ ′ = 0, (5.49)<br />

( ( )<br />

θ θ0<br />

=⇒ tan = tan exp [−(τ − τ 0 )]<br />

2)<br />

2<br />

φ = φ 0 = const.<br />

Bianchi Type V II 0<br />

Non-zero structure constants:<br />

C 2 31 = −C2 13 = C3 12 = −C3 21 = 1,<br />

θ ′ = 0, φ ′ = − cos θ, (5.50)<br />

=⇒ θ = θ 0 = const.<br />

φ = φ 0 − (τ − τ 0 ) cos θ 0 .


5.7 PATTERNS OF ANISOTROPIC MODELS ON THE <strong>CMB</strong> SKY 89<br />

Bianchi Type VII h<br />

Non-zero structure constants:<br />

C 2 31 = −C2 13 = C3 12 = −C3 21 = 1,<br />

C 2 12 = −C 2 21 = C 3 13 = −C 3 31 = √ h.<br />

θ ′ = − √ h sin θ, φ ′ = − cos θ, (5.51)<br />

( ( )<br />

θ θ0<br />

=⇒ tan = tan exp [−(τ − τ 0 )<br />

2)<br />

√ h]<br />

2<br />

φ = φ 0 + (τ − τ 0 ) − √ 1 ( ) ( )<br />

ln {sin 2 θ0<br />

+ cos 2 θ0<br />

exp [2(τ − τ 0 ) √ h]}.<br />

h 2<br />

2<br />

Bianchi Type IX<br />

C A BC = ɛ ABC<br />

θ ′ = φ ′ = 0 (5.52)<br />

=⇒ θ = θ 0 = const.<br />

φ = φ 0 = const.<br />

5.7 Patterns of Anisotropic Models on the <strong>CMB</strong><br />

Sky<br />

In the previous section, we found the null geodesic solutions in Bianchi types that<br />

include FLRW model as an isotropic limit. Now we will compute the effect of these<br />

models on the mean temperature of <strong>CMB</strong> sky. In this section, we do our calculations<br />

<strong>with</strong> the assumption of small anisotropy which means<br />

β ij ≪ 1, |U i | ≪ 1, |σ ij | ≃ |∂ t β ij | ≪ ∂ t α. (5.53)<br />

Relative to an observer <strong>with</strong> 4-velocity U µ , the null vector K µ determines a redshift<br />

factor (−K µ U µ ). Hence the observed cosmological redshift z for comoving matter is<br />

given by Ellis & van Elst (1999)<br />

1 + z E = (K µU µ ) E<br />

(K µ U µ ) 0<br />

(5.54)


5.7 PATTERNS OF ANISOTROPIC MODELS ON THE <strong>CMB</strong> SKY 90<br />

where subscript E and O refer to photon emission and reception respectively, U µ 0<br />

is the<br />

velocity vector of the observer, U µ E<br />

is the velocity vector of the emitting surface. We want<br />

to study the effect of the redshift given in eqn. (5.54) on cosmic microwave background<br />

radiation which is considered to have a black body spectrum at a temperature T E ,<br />

coming from the surface of last scattering. The observed temperature T 0 is then<br />

T 0 =<br />

T E<br />

1 + z E (θ 0 , φ 0 )<br />

(5.55)<br />

If we define p i to be the direction cosines of the null geodesic, p i = K i /K = (cos θ, sin θ cos φ, sin θ sin φ)<br />

the above expression to first order, will become<br />

T 0 = T E Γ −1 [ 1 + (p i U i ) 0 − (p i U i ) E − Σ ] . (5.56)<br />

where<br />

Γ = e α E−α 0<br />

, Σ =<br />

∫ 0<br />

E<br />

p i p k σ ik dt<br />

The <strong>CMB</strong> pattern in eqn. (5.56) has an easy interpretation. The first term on the<br />

right hand side is a Doppler term and exhibits a dipole in the temperature anisotropy<br />

due to the motion of the observer relative to the hypersurfaces of constant time at the<br />

time of observation. This causes a net dipole. The second term represents a Doppler<br />

shift due to the peculiar motion of the source of emitting photon on the surface of last<br />

scattering. This will cause a local redshift/blueshift but usually not a net dipole. In<br />

type I and IX, p i E<br />

= pi 0 , so the second term also gives a dipole variation Collins &<br />

Hawking (1973b). These two terms are absent in orthogonal models but are present in<br />

tilted models. The third term, Σ, gives the temperature anisotropy cause by the shear<br />

of the Universe due to the anisotropic expansion. In types I and IX this causes a pure<br />

quadrupole pattern equal to p i 0 pj 0 (β 0 − β E ) ij .<br />

To obtain the signature of Bianchi models on <strong>CMB</strong> sky, we should solve equation<br />

(5.71). The time dependence of U i and σ ij are given by Einstein equations. And the<br />

variation of θ and φ <strong>with</strong> time obtained from the (null-) geodesic equation which was<br />

discussed in section 5.6. Following is a very brief review of this matter.


5.7 PATTERNS OF ANISOTROPIC MODELS ON THE <strong>CMB</strong> SKY 91<br />

5.7.1 <strong>CMB</strong> Patterns in Type I Bianchi Models<br />

This is the simplest type for which the null geodesic direction cosines are all constant,<br />

π i = constant. α and β are given by<br />

e α = t 2/3 , β = 0 (5.57)<br />

The peculiar velocity is zero and the vorticity vanishes. For small perturbations, β ij =<br />

A ij t −1 and the shear is σ ij = −A ij t −2 where A ij is a constant matrix. Hence the only<br />

contribution to <strong>CMB</strong> pattern is a quadrupole which comes from the Σ term and is equal<br />

to<br />

p i 0 pj 0 A ij(t −1<br />

0 − t −1<br />

E ) (5.58)<br />

5.7.2 <strong>CMB</strong> Patterns in Type V Bianchi Models<br />

This models is the simplest generalization of the k = −1 FLRW models. α and β<br />

are given by<br />

e α = 8πM<br />

3<br />

sinh 2 τ/2, β = 0 (5.59)<br />

where τ is a new time variable defined by 6 dt/dτ = e α <strong>with</strong> τ = 0 when e α = 0, and<br />

M = ρ 0 (H0 2 − 8πρ 0/3) −3/2 ρ<br />

=<br />

0<br />

. High density type V models behave like type<br />

H0 3(1−Ω 0) 3/2<br />

I models, so here we consider low density models. The null geodesics for these models<br />

are calculated in the previous section. For small anisotropies, σ ij = A ij e −3α where A ij<br />

is a constant matrix given by A 1j = (8πM/3)U j e α and A 23 = 0 (one can choose the 2-3<br />

axis such that this happens). The Σ term will then cause an anisotropy in this form<br />

− 3 2 A 11I 1 + 1 2 (A 22 − A 33 )I 1 cos φ + 2(A 12 cos φ + A 13 sin φ)I 2 (5.60)<br />

where<br />

I 1 =<br />

∫ 0<br />

E<br />

e −2α cos 2 θdτ, and I 2 =<br />

∫ 0<br />

E<br />

e −2α sin θ cos θdτ. (5.61)<br />

6 I will denote ∂ t α 0 which is the present value of the expansion rate as H 0 .


5.7 PATTERNS OF ANISOTROPIC MODELS ON THE <strong>CMB</strong> SKY 92<br />

5.7.3 <strong>CMB</strong> Patterns in Type VII 0 Bianchi Models<br />

This is the most general family of models which include the k = 0 FLRW solution.<br />

α and β are given by<br />

e α = ( 9t2<br />

4H 0<br />

) 1/3 x, β = 0 (5.62)<br />

where x is a constant. The parameter x is equal to the characteristic length scale, e α 0<br />

,<br />

over which the invariant vectors change their orientation, multiplied to the expansion<br />

rate, H 0 . The value of parameter x is does not make any physical difference to the FLRW<br />

model but it governs the length scale of the anisotropy in an anisotropic model.For<br />

small anisotropies, β is not zero as given in Collins & Hawking (1973b). Null geodesic<br />

equations are calculated in the previous section. The Doppler terms (p i U i ) 0 and (p i U i ) E<br />

arising from the 1-component of the peculiar velocity give rise to a dipole anisotropy<br />

of magnitude U (1)<br />

0 (Γ − 1) cos θ 0 . The 2 and 3 components of the peculiar velocity will<br />

cause the following pattern<br />

( [ ( 2<br />

Γ cos<br />

x<br />

(<br />

− Γ sin<br />

1 − 1<br />

Γ 1/2 )<br />

]<br />

cos θ<br />

[ 2<br />

x<br />

(<br />

1 − 1<br />

Γ 1/2 )<br />

cos θ<br />

])<br />

)<br />

− 1<br />

(U (2)<br />

0 cos φ + U (3)<br />

0 sin φ) sin θ (5.63)<br />

(U (2)<br />

0 sin φ − U (3)<br />

0 cos φ) sin θ<br />

For x ≪ 1 this is a tightly wound spiral pattern which performs 2/pix(1 − 1/Γ 1/2 )<br />

revolutions and has spacing πx/ sin θ 0 [Γ 1/2 /(Γ 1/2 −1)] between successive maxima Collins<br />

& Hawking (1973b). The Σ term gives a quadrupole pattern through σ 11 which is like<br />

(β 0 − β E ) 11 cos 2 θ. The effect of the other components of σ ij is complicated and some<br />

approximate results near the equator has been given in Collins & Hawking (1973b).<br />

For x ≫ 1<br />

the second term vanishes and it becomes just a dipole pattern (Γ −<br />

1)((U (2)<br />

0 cos φ + U (3)<br />

0 sin φ) sin θ. Also the p i 0<br />

gives a quadrupole pattern p i p j (β 0 − β E ) ij .<br />

are almost constant and hence the Σ term<br />

5.7.4 <strong>CMB</strong> Patterns in Type VII h Bianchi Models<br />

This is the most general family of models which includes the k = 0 FLRW solutions.<br />

It has some of the features of both types V and type VII 0 . There is an adjustable


5.7 PATTERNS OF ANISOTROPIC MODELS ON THE <strong>CMB</strong> SKY 93<br />

parameter h in this model which is given by the square of structure constants (see<br />

previous section). In an unperturbed FLRW model, the expansion scale factor, α, and<br />

β are given by<br />

e α =<br />

As in type VII 0 , we introduce a factor x = H 0 e α 0<br />

by<br />

( )<br />

h 1/2 Ω 0<br />

h 1/2 τ<br />

H 0 (1 − Ω 0 ) 3/2 sinh2 , β = 0 (5.64)<br />

2<br />

which is related to the parameter h<br />

√<br />

h<br />

x =<br />

(5.65)<br />

1 − Ω 0<br />

As we said before, x has no physical effect on the FLRW models. It can be seen from<br />

the fact that the present value of the scale factor can be arbitrarily chosen, but from<br />

(5.64) we have<br />

e α 0<br />

= x H 0<br />

, (5.66)<br />

And hence we see that parameter x can be scaled out of the solution. However, for large<br />

values of x (or equally h), the models are similar to those of type V. As the present<br />

density tends towards the critical density, Ω 0 → 1, and h → 0 in such a way that x<br />

remains finite, the behavior of the models tend to that of type VII 0 . Behavior of these<br />

models <strong>with</strong> different parameters can be seen in figures 5.2 to 5.5.<br />

Null geodesics have been given in the previous section and have a complicated form.<br />

They start near the South Pole and spiral in the negative φ-direction up towards the<br />

equator, and then spiral in the positive φ-direction up towards the North Pole. Small<br />

anisotropies can be treated as perturbations to FLRW model. β has been calculated<br />

for them and is given in Collins & Hawking<br />

(1973b). Here we follow Barrow et al.<br />

(1985) who used these to calculate the contribution to temperature anisotropies from<br />

the vorticity alone. They argue that the inclusion of other pure shear distortions which<br />

are independent of the rotation could only make the temperature anisotropies larger.<br />

And hence their results would give the maximum level of vorticity consistent <strong>with</strong> a<br />

given value of temperature anisotropy.<br />

In type VII h , there are two independent vorticity components ω 2 and ω 3 . They are<br />

given in terms of the off-diagonal shear elements as<br />

ω 2 = (3h − 1)σ 13 − 4h 1/2 σ 12<br />

3x 2 Ω 0<br />

, (5.67)


5.7 PATTERNS OF ANISOTROPIC MODELS ON THE <strong>CMB</strong> SKY 94<br />

ω 3 = (1 − 3h)σ 12 − 4h 1/2 σ 13<br />

3x 2 Ω 0<br />

.<br />

The observables in this model are two dimensionless amplitudes: ( σ 12<br />

H<br />

vorticity is given by<br />

)<br />

and ( σ 13<br />

0 H<br />

)<br />

0 . The<br />

ω = 1 2 e−α (1 + h) 1/2 [ (u 2 ) 2 + (u 3 ) 2] 1/2 , (5.68)<br />

where u 2 and u 3 are the velocity components and their present values are given by<br />

(u 2 ) 0 = 1 [ (<br />

3h 1/2 σ12<br />

3xΩ 0 H<br />

(u 3 ) 0 = 1 [( σ12<br />

3xΩ 0 H<br />

Hence the present value of vorticity is<br />

( ω<br />

H<br />

)<br />

0<br />

)<br />

0<br />

)<br />

0<br />

−<br />

( σ13<br />

H<br />

+ 3h 1/2 ( σ13<br />

H<br />

= (1 + [ h)1/2 (1 + 9h) 1/2 (σ12 ) 2<br />

+<br />

6x 2 Ω 0 H 0<br />

To first order, temperature fluctuations are given by<br />

) ]<br />

, (5.69)<br />

0<br />

) ]<br />

.<br />

0<br />

( σ13<br />

) 2<br />

0<br />

H<br />

] 1/2<br />

. (5.70)<br />

∆T (θ 0 , φ 0 )<br />

T 0<br />

≃ (p i U i ) 0 − (p i U i ) E −<br />

∫ 0<br />

E<br />

p i p k σ ik dt, (5.71)<br />

where θ 0 and φ 0 are related to the actual observing angles by<br />

θ = π − θ 0 , φ = π + φ 0 (5.72)<br />

Substituting for null geodesics from eqn. (5.51) and for velocities from eqn. (5.69) into<br />

eqn. (5.71), we will obtain<br />

∆T (θ 0 , φ 0 )<br />

T 0<br />

=<br />

+<br />

[( σ12<br />

)<br />

H<br />

[( σ12<br />

)<br />

H<br />

0<br />

0<br />

A(θ 0 ) +<br />

B(θ 0 ) −<br />

where coefficients A(θ 0 ) and B(θ 0 ) are defined by<br />

( σ13<br />

H<br />

( σ13<br />

H<br />

)<br />

)<br />

0<br />

0<br />

]<br />

B(θ 0 ) sin φ 0 (5.73)<br />

]<br />

A(θ 0 ) cos φ 0 ,<br />

A(θ 0 ) = C 1 [sin θ 0 − C 2 (cos ψ E − 3h 1/2 sin ψ E )] (5.74)<br />

∫ τ0<br />

s(1 − s 2 ) sin φdτ<br />

+ C 3<br />

(1 + s 2 ) 2 sinh 4 h 1/2 τ/2 ,<br />

τ E<br />

B(θ 0 ) = C 1 [3h 1/2 sin θ 0 − C 2 (sin ψ E + 3h 1/2 cos ψ E )]<br />

∫ τ0<br />

s(1 − s 2 ) cos φdτ<br />

− C 3<br />

(1 + s 2 ) 2 sinh 4 h 1/2 τ/2 .<br />

τ E


5.7 PATTERNS OF ANISOTROPIC MODELS ON THE <strong>CMB</strong> SKY 95<br />

The limits of integration are defined as<br />

τ 0 = 2h −1/2 sinh −1 (Ω −1<br />

0 − 1) 1/2 , (5.75)<br />

τ E = 2h −1/2 sinh −1 ( Ω<br />

−1<br />

0 − 1<br />

1 + z E<br />

) 1/2<br />

,<br />

and constants C 1 , C 2 , C 3 , s and ψ are defined by<br />

C 1 = (3Ω 0 x) −1 ; (5.76)<br />

C 2 = 2s E(1 + z E )<br />

;<br />

1 + s 2 E<br />

C 3 = 4h 1/2 (1 − Ω 0 ) 3/2 Ω<br />

( ( −2<br />

0<br />

)<br />

;<br />

θ θ0<br />

s = tan = tan exp [−(τ − τ 0 )<br />

2)<br />

√ h],<br />

2<br />

ψ = φ 0 + (τ − τ 0 ) − √ 1 ( ) ( )<br />

ln {sin 2 θ0<br />

+ cos 2 θ0<br />

exp [2(τ − τ 0 ) √ h]}.<br />

h 2<br />

2<br />

The expression for ∆T (θ 0,φ 0 )<br />

T 0<br />

where ˜φ is defined as<br />

and σ is given by<br />

can be written in a compact form<br />

∆T<br />

T 0<br />

(θ 0 , φ 0 ) = (A 2 + B 2 ) 1/2 ( σ<br />

H<br />

cos ˜φ =<br />

[( σ12<br />

σ<br />

)<br />

B −<br />

( σ13<br />

σ<br />

)<br />

0<br />

cos (φ 0 + ˜φ) (5.77)<br />

) ]<br />

A (A 2 + B 2 ) −1/2 (5.78)<br />

σ 2 = σ 2 12 + σ 2 13 (5.79)<br />

Eqn. (5.77) helps to understand the behavior of the <strong>CMB</strong> pattern in these models. If<br />

we look around any circle at a given θ 0 on the sky, the temperature variation will have<br />

a pure cos φ 0 behavior. ˜φ determines the relative orientation of adjacent θ0 =constant<br />

rings. In type V models, A(θ 0 ) = 0 and hence we have no spiraling in these models. On<br />

the other hand, B(θ 0 ) determines the amplitude of ∆T<br />

T 0<br />

and the focusing into a hot spot<br />

[Barrow et al. (1985)]. Some of these maps have been shown in figures 5.2 to 5.5.<br />

5.7.5 <strong>CMB</strong> Patterns in Type IX Bianchi Models<br />

This is a generalization of the k = +1 FLRW models <strong>with</strong><br />

e α = s α 0<br />

sin τ/4 , β = 0 (5.80)


5.8 UNVEILING HIDDEN PATTERNS OF BIANCHI VII H 96<br />

Figure 5.2 Bianchi patterns in type VII h models for x = 0.1 and Ω 0 = 0.1 (left),Ω 0 = 0.5<br />

(middle) and Ω 0 = 0.9 (right).<br />

For small anisotropies, β ij is given by a perturbative series derived by Collins & Hawking<br />

(1973b). The null geodesics have p i = constant and hence the Doppler terms (p i U i ) 0<br />

and (p i U i ) E give rise to a dipole anisotropy equal to (p i U i ) 0 (Γ − 1). The Σ term causes<br />

a quadrupole anisotropy of the form p i p j (β 0 − β E ) ij .<br />

5.8 Unveiling Hidden Patterns of Bianchi VII h<br />

Recently Bianchi VII h models were compared to the first-year WMAP data on large<br />

scales and it was shown that the best fit Bianchi model corresponds to a highly hyperbolic<br />

model (Ω 0 = 0.5) <strong>with</strong> a right-handed vorticity ( ω ) H 0 = 4.3 × 10 −10 [Jaffe et<br />

al. (2005)]. There are several aspects to this hidden pattern which should be mentioned.<br />

First of all, the proposed model is an extremely hyperbolic model and does not<br />

agree <strong>with</strong> the location of the first peak in the best fit power spectrum of the <strong>CMB</strong>. In<br />

the regime of VII h models, one should study the nearly flat models in order to have a<br />

consistent angular power spectrum. But the problem is that as it has been shown in<br />

figs 5.2 to 5.5, the spiral patterns in high density models are not concentrated in one<br />

hemisphere and can’t be used to cure the large-scale power asymmetry in the WMAP<br />

data reported by Eriksen et al. (2004a). Second is that the Bianchi patterns we studied<br />

so far, are only valid for a matter dominated universe. Patterns must be recalculated in<br />

a universe <strong>with</strong> a dark energy component if one wants to compare them <strong>with</strong> the real<br />

observed <strong>CMB</strong> data which is out there in a Λ dominated universe.<br />

For these reasons, we study the type VII h Bianchi models only as hidden patterns<br />

in the <strong>CMB</strong> anisotropy maps. We pay our attention to the specific model proposed by


5.8 UNVEILING HIDDEN PATTERNS OF BIANCHI VII H 97<br />

Jaffe et al. (2005) and in the next two sections, we address the question whether and<br />

to what extent one would be able to discover this pattern (Fig. 5.7) and other hidden<br />

anisotropic pattern in the <strong>CMB</strong>.


5.8 UNVEILING HIDDEN PATTERNS OF BIANCHI VII H 98<br />

Figure 5.3 Bianchi patterns in type VII h models for x = 0.55 and Ω 0 = 0.1 (left),Ω 0 = 0.5<br />

(middle) and Ω 0 = 0.9 (right).<br />

Figure 5.4 Bianchi patterns in type VII h models for x = 1 and Ω 0 = 0.1 (left),Ω 0 = 0.5<br />

(middle) and Ω 0 = 0.9 (right).<br />

Figure 5.5 Bianchi patterns in type VII h models for x = 5 and Ω 0 = 0.1 (left),Ω 0 = 0.5<br />

(middle) and Ω 0 = 0.9 (right).


5.8 UNVEILING HIDDEN PATTERNS OF BIANCHI VII H 99<br />

1500<br />

C l l (l+1) / 2 π (µK) 2<br />

1000<br />

500<br />

0<br />

0 10 20 30<br />

Multipole l<br />

Figure 5.6 Top panel: WMAP Internal Linear Combination map. Middle panel: Best-fit<br />

Bianchi VII h template (enhanced by a factor of four to show the structure). Bottom<br />

panel: Difference between ILC and best-fit Bianchi template; the “Bianchi-corrected”<br />

ILC map. Figure and caption courtesy of Jaffe et al. (2005).<br />

Figure 5.7 Comparison of power spectra. The gray and black solid lines show the power<br />

spectrum estimated from the co-added V+W map before and after correcting for the<br />

Bianchi template, respectively. The dotted gray and black lines shows the theoretical<br />

best-fit power-spectra from the WMAP-team analysis and Hansen et al. (2004) respectively.<br />

The latter is a fit to northern hemisphere data alone. The light gray dashed line<br />

corresponds to the full sky power spectrum of the best-fit Bianchi template. Figure and<br />

caption from Jaffe et al. (2005).


5.8 UNVEILING HIDDEN PATTERNS OF BIANCHI VII H 100<br />

Bipolar Power Spectrum Analysis<br />

To do a statistical study of these patterns, we generate the pattern for a given model.<br />

This pattern is given by the shear components, x parameter and the Ω 0 . We will then<br />

simulate random <strong>CMB</strong> maps from the best fit C l of the WMAP data. We add these<br />

two maps <strong>with</strong> a strength factor α which will let us control the relative strength of the<br />

pattern and the random map (see Fig. 5.8). The resultant map is then given by<br />

Figure 5.8 Adding a pattern template <strong>with</strong> a strength α to a random realization of the<br />

<strong>CMB</strong> anisotropy map. Maps are rotated to the Galactic center for illustration<br />

∆T (ˆn) = ∆T <strong>CMB</strong> (ˆn) + α∆T Bianchi (ˆn), (5.81)<br />

and hence the power spectrum of this map will be given by<br />

C l = C <strong>CMB</strong><br />

l<br />

+ α 2 C Bianchi<br />

l (5.82)<br />

α is related to ( σ ) H 0 through ( σ ) H 0 = α T 0<br />

where T 0 = 2.73 × 10 6 µK.<br />

We analysis the above Bianchi added <strong>CMB</strong> maps using BiPS method. Since BiPS is<br />

orientation independent, we don’t need to worry about the relative orientation between<br />

the background map and the Bianchi template. The procedure is then as follows. For<br />

different strength factors we simulate 100 Bianchi added <strong>CMB</strong> maps 7<br />

at HEALPix<br />

resolution of N side = 32 corresponding to l max = 95. Some of these maps are shown<br />

in Figure 5.9.<br />

eqn. (5.81) are a lm = a <strong>CMB</strong><br />

lm<br />

BiPS of each map is obtained from the total a lm which according to<br />

+ α a Bianchi<br />

lm . We compute the BiPS for each map using the<br />

estimator given in eqn. (2.37). We average these 100 BiPS and compute the dispersion<br />

in them. The dispersion is an estimate of 1σ error bars. At last we use the total<br />

angular power spectrum given in eqn. (5.82) to estimate the bias. We use the best fit<br />

7 Simulation of background maps and the spherical harmonic expansion of the resultant map is done<br />

by HEALPix.


5.8 UNVEILING HIDDEN PATTERNS OF BIANCHI VII H 101<br />

Figure 5.9 Four Bianchi added <strong>CMB</strong> maps <strong>with</strong> different strength factors, rotated to<br />

the Galactic center for illustration. The Bianchi template in the above maps has been<br />

computed <strong>with</strong> x = 0.55, Ω 0 = 0.5 and strength factors ( σ ) H 0 = 1.83 × 10 −9 (top left),<br />

( σ ) H 0 = 1.09 × 10 −9 (top right), ( σ ) H 0 = 7.3 × 10 −10 (bottom left), ( σ ) H 0 = 3.66 × 10 −10<br />

(bottom right).<br />

theoretical power spectrum from the WMAP analysis, Spergel et al.<br />

(2003), as C <strong>CMB</strong><br />

l<br />

in computing the bias by the estimator of eqn. (2.40). Some results of BiPS for different<br />

strength factors are shown in Figure 5.10. As it has been discussed in Chapter 2, we<br />

can use different window functions in harmonic space, W l , in order to concentrate on<br />

a particular l-range. This is proved to be a useful and strong tool to detect deviations<br />

from SI using BiPS method. In other words, multipole space windows that weigh down<br />

the contribution from the SI region of multipole space will enhance the signal relative<br />

to cosmic error. We use simple filter functions in l space to isolate different ranges of<br />

angular scales; a low pass, Gaussian filter, Wl<br />

G (l s ) and a band pass filter, Wl S (l t , l s ).<br />

These filters are plotted in the lower panel of Figure 7.1.<br />

We use a simple χ 2 statistics to compare our BiPS results <strong>with</strong> zero,<br />

l∑<br />

max<br />

χ 2 =<br />

l=0<br />

(<br />

κl<br />

σ κl<br />

) 2<br />

. (5.83)


5.8 UNVEILING HIDDEN PATTERNS OF BIANCHI VII H 102<br />

0<br />

0<br />

5 10 15<br />

5 10 15<br />

0<br />

0<br />

5 10 15<br />

5 10 15<br />

Figure 5.10 Bipolar power spectrum of Bianchi added <strong>CMB</strong> maps <strong>with</strong> strength factors<br />

of ( σ H ) 0 = 4.76 × 10 −10 , 3.66 × 10 −10 , 2.56 × 10 −10 , 1.25 × 10 −10 . We have used W S<br />

l (4, 13)<br />

to filter the maps.<br />

The probability of detecting a map <strong>with</strong> a given BiPS is then given by the probability<br />

distribution of the above χ 2 . Probability of χ 2 versus the strength factor is plotted in<br />

Figure 5.11. Being on the conservative side we can constrain the strength factor to be<br />

( σ ) H 0 = 3.15 × 10 −10 at a 99.9% confidence level, Ghosh et al. (2006).<br />

We can repeat the above exercise for nearly flat (Ω 0 = 0.99) Bianchi VII h templates.<br />

Some of these Bianchi added <strong>CMB</strong> maps are shown in Figure 5.12. As we expect, the<br />

spiral patterns of these models are wider than the hyperbolic models and is not focused<br />

in one hemisphere. As it can be seen in Figure 5.13 they have a weaker BiPS signal.<br />

And hence we get a weaker limit of ( σ ) H 0 = 4.7 × 10 −10 at a 99.9% confidence level on<br />

them (see Figure 5.14).


5.8 UNVEILING HIDDEN PATTERNS OF BIANCHI VII H 103<br />

5<br />

1<br />

0.32<br />

0.1<br />

P(χ 2 )<br />

0.01<br />

0.001<br />

0.0001<br />

2.52<br />

2.83<br />

3.15<br />

3.48<br />

4.3<br />

4.7<br />

(σ/H) 0 in (10 -10 )<br />

Figure 5.11 Probability distribution of χ 2 versus the strength factor for Ω 0 = 0.5 Bianchi<br />

template. Results of two different window functions, Wl<br />

G (25) (blue line) and Wl S (4, 13)<br />

(green line) are shown.<br />

Figure 5.12 Two Bianchi added <strong>CMB</strong> maps, Ω 0 = 0.99, <strong>with</strong> different strength factors,<br />

( σ H ) 0 = 1.09 × 10 −9 and 3.66 × 10 −10 , rotated to the Galactic center for illustration.


5.8 UNVEILING HIDDEN PATTERNS OF BIANCHI VII H 104<br />

0<br />

0<br />

5 10 15<br />

5 10 15<br />

Figure 5.13 Bipolar power spectrum of Bianchi added <strong>CMB</strong> maps <strong>with</strong> strength factors<br />

of ( σ H ) 0 = 7.33 × 10 −10 (left) and ( σ H ) 0 = 4.76 × 10 −10 (right).<br />

5<br />

1<br />

0.32<br />

0.1<br />

P(χ 2 )<br />

0.01<br />

0.001<br />

0.0001<br />

3.96 4.4 4.7<br />

(σ/H) 0 in (10 -10 )<br />

5 5.55.8<br />

Figure 5.14 Probability distribution of χ 2 versus the strength factor for Ω 0 = 0.99<br />

Bianchi template. Results of two different window functions, Wl<br />

G (25) (red line) and<br />

W S<br />

l (4, 13) (cyan line) are shown.


Chapter 6<br />

Observational Artifacts<br />

Foregrounds and observational artifacts (such as non-circular beam, incomplete/nonuniform<br />

sky coverage and anisotropic noise) would also manifest themselves as violations<br />

of SI.<br />

6.1 Anisotropic noise<br />

The <strong>CMB</strong> temperature measured by an instrument is a linear sum of the cosmological<br />

signal as well as instrumental noise. The two point correlation function then has two<br />

parts, one part comes from the signal and the other one comes from the noise<br />

C(ˆn 1 , ˆn 2 ) = C S (ˆn 1 , ˆn 2 ) + C N (ˆn 1 , ˆn 2 ). (6.1)<br />

Both signal and noise should be statistically isotropic to have a statistically isotropic<br />

<strong>CMB</strong> map. So even for a statistically isotropic signal, if the noise fails to be statistically<br />

isotropic the resultant map will turn out to be anisotropic. The noise matrix can fail<br />

to be statistically isotropic due to non-uniform coverage. Also if the noise is correlated<br />

between different pixels the noise matrix could be statistically anisotropic. A simple<br />

example of this is the diagonal (but anisotropic) noise given by the following correlation<br />

C N (ˆn, ˆn ′ ) = σ 2 (ˆn)δˆnˆn ′. (6.2)<br />

105


6.1 ANISOTROPIC NOISE 106<br />

This noise clearly violates the SI and will result a non-zero BiPS given by<br />

κ l =<br />

l∑<br />

m=−l<br />

|f lm | 2 , (6.3)<br />

where f lm are spherical harmonic transform of the noise, f lm = ∫ dΩˆn Y ∗<br />

lm (ˆn)σ2 (ˆn).<br />

As an example we study the BiPS of the coadded noise covariance matrix of WMAP<br />

which is a simple diagonal matrix. We have plotted this matrix in figure 6.1 where the<br />

value on each pixel represents the 1/σ 2 i<br />

on that pixel. We compute the BiPS prediction<br />

Figure 6.1 The coadded noise covariance matrix of WMAP. This is in fact 1/σi 2 . The<br />

two spots represent cleaner regions <strong>with</strong> less noise.<br />

for this case using eqn. (6.3). The result is shown in figure 6.2. The characteristic<br />

component of this covariance matrix is the l = 2 component. This is very well expected<br />

and can be explained. This is a two point correlation of noise <strong>with</strong> a preferred direction.<br />

As it is shown in Chapter 3, the signature of such a preferred axis is a non-zero l = 2<br />

component in BiPS. It is always very important to know how well a statistical measure<br />

can be estimated from a map. We would like to know what would we see if we estimate<br />

BiPS from a <strong>CMB</strong> map. To answer that one should make realizations of the above<br />

covariance matrix and add them to simulated <strong>CMB</strong> maps and the estimate the BiPS<br />

from them and study the average BiPS obtained from many realizations. However at


6.2 ANISOTROPIC NOISE 107<br />

4e-05<br />

"bips_cttp_coadd_noise_3.5deg_32.dat"<br />

3.5e-05<br />

3e-05<br />

2.5e-05<br />

2e-05<br />

1.5e-05<br />

1e-05<br />

5e-06<br />

0<br />

0 2 4 6 8 10 12 14<br />

Figure 6.2 BiPS prediction for the diagonal noise covariance matrix of Figure 6.1. The<br />

characteristic component is the l = 2.<br />

this point we are only interested in the behavior of the noise patterns. So, we will stick<br />

to estimating BiPS from noise maps solely. We simulate noise maps from the noise<br />

covariance matrix of Fig. 6.1. To do that we use our fast methods of generating <strong>CMB</strong><br />

maps from anisotropic correlation functions explained in Appendix F. The result is<br />

shown in Fig. F.1 and we see that the noise dispersion is smaller on the place of those<br />

two clean spots of Fig. 6.1. We make 100 realizations of noise maps and compute the<br />

BiPS for them. The average BiPS has a clear non-zero component at l = 2. Fig. 6.3<br />

shows this for simulated noise maps.


6.2 THE EFFECT OF NON-CIRCULAR BEAM 108<br />

0<br />

0<br />

2 4 6 8<br />

Figure 6.3 Estimating BiPS from 100 simulations of noise maps <strong>with</strong> a sharp cut off in<br />

l-space (top) and <strong>with</strong> a Gaussian filter, G(120), (bottom).<br />

6.2 The effect of non-circular beam<br />

as<br />

The temperature of the <strong>CMB</strong> anisotropy measured by an instrument can be written<br />

∫<br />

∆ ˜T (ˆn) = ∆T (ˆn ′ )B(ˆn, ˆn ′ )dΩˆn ′. (6.4)<br />

Here B(ˆn, ˆn ′ ) is the beam function that encodes the instrumental response, ˆn denotes the<br />

direction to the center of the beam and ˆn ′ denotes the direction of the incoming photon.<br />

Using this relation to calculate the correlation function ˜C(ˆn 1 , ˆn 2 ) = 〈∆ ˜T (ˆn 1 )∆ ˜T (ˆn 2 )〉<br />

one would get<br />

∫ ∫<br />

˜C(ˆn 1 , ˆn 2 ) = dΩˆn ′ dΩˆn ′′〈∆T (ˆn ′ )∆T (ˆn ′′ )〉B(ˆn 1 , ˆn ′ )B(ˆn 2 , ˆn ′′ ) (6.5)


6.3 MASK EFFECTS 109<br />

=<br />

∫<br />

dΩˆn ′<br />

∫<br />

dΩˆn ′′C(ˆn ′ , ˆn ′′ )B(ˆn 1 , ˆn ′ )B(ˆn 2 , ˆn ′′ ).<br />

Only for a circular beam where B(ˆn, ˆn ′ ) ≡ B(ˆn · ˆn ′ ), the correlation function is statistically<br />

isotropic, ˜C(ˆn1 , ˆn 2 ) ≡ ˜C(ˆn 1 · ˆn 2 ). Breakdown of SI is obvious since even C l get<br />

mixed for a non-circular beam, ˜Cl = ∑ l<br />

A ′ ll ′C l ′ Mitra et al. (2004). Non-circularity<br />

of the beam in <strong>CMB</strong> anisotropy experiments is becoming increasingly important as<br />

experiments go for higher resolution measurements at higher sensitivity.<br />

6.3 Mask effects<br />

Most <strong>CMB</strong> experiments map only a part of the sky. Even in the best case, contamination<br />

by galactic foreground residuals make parts of the sky unusable. The incomplete<br />

sky or mask effect is another source of breakdown of SI. But, this effect can be readily<br />

modeled out. The effect of a general mask on the temperature field is as follows<br />

∆T masked (ˆn) = ∆T (ˆn)W (ˆn), (6.6)<br />

where W (ˆn) is the mask function. One can cut different parts of the sky by choosing<br />

appropriate mask functions. Masked a lm coefficients can be computed from the masked<br />

temperature field,<br />

a masked<br />

lm =<br />

∫<br />

∆T masked (ˆn)Y ∗<br />

lm (ˆn)dΩˆn (6.7)<br />

= ∑ l 1 m 1<br />

a l1 m 1<br />

∫<br />

Y l1 m 1<br />

(ˆn)Y ∗<br />

lm(ˆn)W (ˆn)dΩˆn .<br />

Where a l1 m 1<br />

are spherical harmonic transforms of the original temperature field. We<br />

can expand W (ˆn) in spherical harmonics as well<br />

W (ˆn) = ∑ lm<br />

w lm Y lm (ˆn), (6.8)<br />

and after substituting this into eqn. (6.7) it is seen that the masked a lm is given by the<br />

effect of a kernel K l 1m 1<br />

lm on original a lm Prunet et al. (2004)<br />

a masked<br />

lm = ∑ l 1 m 1<br />

a l1 m 1<br />

K l 1m 1<br />

lm . (6.9)


6.4 RESIDUALS FROM FOREGROUND REMOVAL 110<br />

The kernel contains the information of our mask function and is defined by<br />

K l 1m 1<br />

lm<br />

= ∑ ∫<br />

w l2 m 2<br />

Y l1 m 1<br />

(ˆn)Y l2 m 2<br />

(ˆn)Ylm ∗ (ˆn)dΩˆn (6.10)<br />

l 2 m 2<br />

= ∑ l 2 m 2<br />

w l2 m 2<br />

√<br />

(2l 1 + 1)(2l 2 + 1)<br />

Cl l0<br />

4π(2l + 1) 1 0l 2 0Cl lm<br />

1 m 1 l 2 m 2<br />

.<br />

In the last step of the above expression we used rule G.3 of spherical harmonics to<br />

simplify the expression. The covariance matrix of a masked sky will no longer have the<br />

diagonal form of eqn.(2.17) because of the action of the kernel<br />

〈a masked<br />

lm a masked l ∗<br />

′ m 〉 = 〈a ′ l 1 m 1<br />

a ∗ l 1 ′ m′ 〉Kl 1m 1<br />

1 lm Kl′ 1 m′ 1<br />

l ′ m<br />

(6.11)<br />

′<br />

= C l1 δ l1 l 1 ′ δ m 1 m ′ Kl 1m 1<br />

1 lm Kl′ 1 m′ 1<br />

l ′ m ′<br />

= ∑<br />

C l1 K l 1m 1<br />

lm Kl 1m 1<br />

l ′ m<br />

. ′<br />

l 1 ,m 1<br />

This clearly violates the SI and results a non-zero BiPS for masked <strong>CMB</strong> skies. In the<br />

next section we apply a galactic mask to ILC map and show that signature of this mask<br />

on BiPS is a rising tail at low l, (l < 20).<br />

6.4 Residuals from foreground removal<br />

Besides the cosmological signal and instrumental noise, a <strong>CMB</strong> map also contains<br />

foreground emission such as galactic emission, etc. The foreground is usually modeled<br />

out using spectral information. However, residuals from foreground subtractions in the<br />

<strong>CMB</strong> map will violate SI. Interestingly, BiPS does sense the difference between maps<br />

<strong>with</strong> grossly different emphasis on the galactic foreground. As an example we compute<br />

BiPS of a Wiener filtered map in the next section to make this point clear (see Fig. 7.4).<br />

It shows a signal very similar to that of a galactic cut sky. This can be understood if<br />

one writes the effect of the Wiener filter as a weight on some regions of the map<br />

∆T W (ˆn) = ∆T (ˆn)(1 + W (ˆn)). (6.12)<br />

This explains the similarity between a cut sky and a Wiener filtered map. The effect of<br />

foregrounds on BiPS has recently been studied in great details by Saha et al. (2006).


Chapter 7<br />

Statistical isotropy of <strong>CMB</strong> anisotropy from<br />

WMAP<br />

After the first year of WMAP data, the SI of the <strong>CMB</strong> anisotropy (i.e. rotational<br />

invariance of n-point correlations) has attracted considerable attention. Tantalizing<br />

evidence of SI breakdown (albeit, in very different guises) has mounted in the WMAP<br />

first year sky maps, using a variety of different statistics. It was pointed out that the<br />

suppression of power in the quadrupole and octopole are aligned Tegmark et al. (2004).<br />

Further “multipole-vector” directions associated <strong>with</strong> these multipoles (and some other<br />

low multipoles as well) appear to be anomalously correlated Copi et al. (2004); Schwarz<br />

et al. (2004). There are indications of asymmetry in the power spectrum at low<br />

multipoles in opposite hemispheres Eriksen et al. (2004a); Hansen et al. (2004);<br />

Naselsky et al. (2004). Possibly related, are the results of tests of Gaussianity that<br />

show asymmetry in the amplitude of the measured genus amplitude (at about 2 to 3σ<br />

significance) between the north and south galactic hemispheres Park (2004); Eriksen<br />

et al. (2004b,c). Analysis of the distribution of extrema in WMAP sky maps has<br />

indicated non-Gaussianity, and to some extent, violation of SI Larson & Wandelt (2004).<br />

However, what is missing is a common, well defined, mathematical language to quantify<br />

SI (as distinct from non Gaussianity) and the ability to ascribe statistical significance<br />

to the anomalies unambiguously.<br />

As a statistical tool of searching for departures from SI, we use the bipolar power<br />

spectrum (BiPS) which is sensitive to structures and patterns in the underlying to-<br />

111


7.0 112<br />

tal two-point correlation function Hajian & Souradeep<br />

(2003b); Souradeep & Hajian<br />

(2003). The BiPS is particularly sensitive to real space correlation patterns (preferred<br />

directions, etc.) on characteristic angular scales. In harmonic space, the BiPS at multipole<br />

l sums power in off-diagonal elements of the covariance matrix, 〈a lm a l ′ m ′〉, in the<br />

same way that the ‘angular momentum’ addition of states lm, l ′ m ′ have non-zero overlap<br />

<strong>with</strong> a state <strong>with</strong> angular momentum |l − l ′ | < l < l + l ′ . Signatures, like a lm and<br />

a l+nm being correlated over a significant range l are ideal targets for BiPS. These are<br />

typical of SI violation due to cosmic topology and the predicted BiPS in these models<br />

have a strong spectral signature in the bipolar multipole l space Hajian & Souradeep<br />

(2003a). The orientation independence of BiPS is an advantage since one can obtain<br />

constraints on cosmic topology that do not depend on the unknown specific orientation<br />

of the pattern (e.g., preferred directions).<br />

We carry out measurement of the BiPS, on the following <strong>CMB</strong> anisotropy maps<br />

A) a foreground cleaned map (denoted as ‘TOH’) Tegmark et al. (2004),<br />

B) the Internal Linear Combination map (denoted as ‘ILC’ in the figures) Bennett et<br />

al.<br />

(2003), and<br />

C) a customized linear combination of the QVW maps of WMAP <strong>with</strong> a galactic cut<br />

(denoted as ‘CSSK’).<br />

Also for comparison, we measure the BiPS of<br />

D) a Wiener filtered map of WMAP data (denoted as ‘Wiener’) Tegmark et al.<br />

(2004), and<br />

E) the ILC map <strong>with</strong> a 10 ◦ cut around the equator (denoted as ‘Gal. cut.’).<br />

Angular power spectra of these maps are shown in Fig. 7.1. The best fit theoretical<br />

power spectrum from the WMAP analysis 1<br />

Spergel et al. (2003) is plotted on the same<br />

figure. C l from observed maps are consistent <strong>with</strong> the theoretical curve, C T l<br />

, (except for<br />

low multipoles. The bias and cosmic variance of BiPS depend on the total SI angular<br />

1 Based on an LCDM model <strong>with</strong> a scale-dependent (running) spectral index which best fits the<br />

dataset comprised of WMAP, CBI and ACBAR <strong>CMB</strong> data combined <strong>with</strong> 2dF and Ly-α data


C T l . We use the estimator given in eqn.(2.37) to measure BiPS for the given <strong>CMB</strong> maps.<br />

7.0 113<br />

power spectrum of the signal and noise C l = C S l + C N l . However, we have restricted<br />

our analysis to l < 60 where the errors in the WMAP power spectrum is dominated by<br />

cosmic variance. It is conceivable that the SI violation is limited to particular range of<br />

angular scales. Hence, multipole space windows that weigh down the contribution from<br />

the SI region of multipole space will enhance the signal relative to cosmic error, σ SI<br />

. We<br />

use simple filter functions in l space to isolate different ranges of angular scales; a low<br />

pass, Gaussian filter<br />

W G<br />

l (l s ) = exp(−(l + 1/2) 2 /(l s + 1/2) 2 ) (7.1)<br />

that cuts off power on small angular scales (< 1/l s ) and a band pass filter,<br />

W S<br />

l (l t , l s ) = [2(1 − J 0 ((l + 1/2)/(l t + 1/2)))] exp(−(l + 1/2) 2 /(l s + 1/2) 2 ) (7.2)<br />

that retains power <strong>with</strong>in a range of multipoles set by l t and l s . The windows are<br />

normalized such that ∑ l (l + 1/2)/(l(l + 1))W l = 1, i.e., unit rms for unit flat band<br />

power C l = 1/(l(l+1)). The window functions used in our work are plotted in figure 7.1.<br />

We use the C T l<br />

to generate 1000 simulations of the SI <strong>CMB</strong> maps. a lm ’s are generated<br />

up to an l max of 1024 (corresponding to HEALPix resolution N side = 512). These<br />

are then multiplied by the window functions Wl<br />

G (l s ) and Wl S (l t , l s ). We compute the<br />

BiPS for each realization.<br />

Fig.7.1 shows that the average power spectrum obtained<br />

from the simulation matches the theoretical power spectrum, C T l<br />

realizations. We use C T l<br />

, used to generate the<br />

to analytically compute bias and cosmic variance estimation for<br />

˜κ l . This allows us to rapidly compute BiPS <strong>with</strong> 1σ error bars for different theoretical<br />

We compute the BiPS for all window functions shown in Fig 7.1. Results for three of<br />

these windows are plotted in Figs. 7.2, 7.3, and 7.4. In the low-l regime, where we have<br />

kept the low multipoles, BiPS for all three given maps are consistent <strong>with</strong> zero (Fig.<br />

7.2). But in the intermediate-l regime (Fig. 7.3), although BiPS of ILC and TOH maps<br />

are well consistent <strong>with</strong> zero, the CSSK map shows a rising tail in BiPS due to the<br />

galactic mask. To confirm it, we compute the BiPS for the ILC map <strong>with</strong> a 10-degree<br />

cut around the galactic plane (filtered <strong>with</strong> the same window function). The result is


7.0 114<br />

shown on the top panel of Fig. 7.3. Another interesting effect is seen when we apply a<br />

W S<br />

l (20, 45) filter, where Wiener filtered map has a non zero BiPS very similar to that<br />

of CSSK but weaker (Fig. 7.4). The reason is that Wiener filter takes out more modes<br />

from regions <strong>with</strong> more foregrounds since these are inconsistent <strong>with</strong> the theoretical<br />

model. As a result, a Wiener filtered map at Wl s (20, 45) filter has a BiPS similar to a<br />

cut sky map. The fact that Wiener map has less power at the Galactic plane can even<br />

be seen by eye! Comparing Fig. 7.3 to Fig.7.4 we see that using different filters allows<br />

us to uncover different types of violation of SI in a <strong>CMB</strong> map. In our analysis we have<br />

used a set of filters which enables us to probe SI breakdown on angular scales l < 60.<br />

The BiPS measured from 1000 simulated SI realizations of C T l<br />

is used to estimate the<br />

probability distribution functions (PDF), p(˜κ l ). A sample of the PDF for two windows<br />

is shown in Fig. 7.5. Measured values of BiPS for ILC, TOH and CSSK maps are plotted<br />

on the same plot. BiPS for ILC and TOH maps are located very close to the peak of<br />

the PDF. We compute the individual probabilities of the map being SI for each of the<br />

measured ˜κ l . This probability is obtained by integrating the PDF beyond the measured<br />

˜κ l . To be precise, we compute<br />

P (˜κ l |C T l ) = P (κ l > ˜κ l ) =<br />

= P (κ l < ˜κ l ) =<br />

∫ ∞<br />

˜κ l<br />

dκ l p(κ l ), ˜κ l > 0, (7.3)<br />

∫ ˜κl<br />

−∞<br />

dκ l p(κ l ), ˜κ l < 0.<br />

The probabilities obtained are shown in Figs. 7.6 and 7.8 for W S (20, 30), Wl<br />

G (40) and<br />

W G<br />

l (4). The probabilities for the W S<br />

l (20, 30) window function are greater than 0.25<br />

and the minimum probability at ∼ 0.05 occurs at κ 4 for W G (40). The reason for<br />

systematically lower SI probabilities for W l S (20, 30) as compared to W l G (40) is simply<br />

due to lower cosmic variance of the former. The contribution to the cosmic variance of<br />

BiPS is dominated by the low spherical harmonic multipoles. Filters that suppress the<br />

a lm at low multipoles have a lower cosmic variance.<br />

It is important to note that the above probability is a conditional probability of<br />

measured ˜κ l being SI given the theoretical spectrum C T l<br />

(used to estimate the bias).<br />

A final probability emerges as the Bayesian chain product <strong>with</strong> the probability of the<br />

theoretical C T l<br />

used given data. Hence, small difference in these conditional probabilities<br />

for the two maps are perhaps not necessarily significant.<br />

Since the BiPS is close to


7.0 115<br />

zero, the computation of a probability marginalized over the Cl<br />

T may be possible using<br />

Gaussian (or, improved) approximation to the PDF of κ l . The important role played<br />

by the choice of the theoretical model for the BiPS measurement is shown for a W l that<br />

retains power in the lowest multipoles, l = 2 and l = 3. Assuming Cl T , there are hints<br />

of non-SI detections in the low l’s (left panel of Fig. 7.7). We also compute the BiPS<br />

using a Cl<br />

T for a model that accounts for suppressed quadrupole and octopole in the<br />

WMAP data Shafieloo & Souradeep (2004). The mild detections of a non zero BiPS<br />

vanish for this case (right panel of Fig. 7.7).<br />

In summary, We find no strong evidence for SI violation in the WMAP <strong>CMB</strong><br />

anisotropy maps considered here. We have verified that our null results are consistent<br />

<strong>with</strong> measurements on simulated SI maps. The BiPS measurement reported here<br />

is a Bayesian estimate of the conditional probability of SI (for each κ l of the BiPS)<br />

given an underlying theoretical spectrum Cl T . We point out that the excess power in<br />

the Cl<br />

T <strong>with</strong> respect to the measured C l from WMAP at the lowest multipoles tends to<br />

indicate mild deviations from SI. BiPS measurements are shown to be consistent <strong>with</strong><br />

SI assuming an alternate model Cl<br />

T that is consistent <strong>with</strong> suppressed power on low<br />

multipoles.


7.0 116<br />

Figure 7.1 Top: C l of the two WMAP <strong>CMB</strong> anisotropy maps. The red, magenta and<br />

green curves correspond to map A, B and C, respectively.<br />

The black line is a ‘best<br />

fit’ WMAP theoretical C l used for simulating SI maps. Blue dots are the average<br />

C l recovered from 1000 realizations. Bottom: These plots show the window functions<br />

used. The dashed curves <strong>with</strong> increasing l coverage are ‘low-pass’ filter, Wl<br />

G (l s ), <strong>with</strong><br />

l s = 4, 18, 40, respectively. The solid lines are ‘band-pass’ filter W S<br />

l (l t , l s ) <strong>with</strong> (l s , l t ) =<br />

(13, 2), (30, 5), (30, 20), (45, 20), respectively.


7.0 117<br />

100<br />

0<br />

-100<br />

-200<br />

CSSK<br />

100<br />

0<br />

-100<br />

-200<br />

ILC<br />

100<br />

0<br />

-100<br />

-200<br />

TOH<br />

5 10 15 20<br />

Figure 7.2 Measured values of κ l for maps A, B and C filtered <strong>with</strong> a Gaussian window<br />

<strong>with</strong> l s = 40. Each map appears to show the violation of statistical isotropy for some<br />

l at the level of ∼ 1σ. However, this is because the PDF for these models are skewed<br />

<strong>with</strong> the maximum probability shifted away from the mean to negative values (see Fig.<br />

7.5).


7.0 118<br />

20<br />

10<br />

0<br />

-10<br />

Gal. cut<br />

20<br />

10<br />

0<br />

-10<br />

Wiener<br />

20<br />

10<br />

0<br />

-10<br />

CSSK<br />

20<br />

10<br />

0<br />

-10<br />

ILC<br />

20<br />

10<br />

0<br />

-10<br />

TOH<br />

5 10 15 20<br />

Figure 7.3 Measured BiPS for maps A, B and C filtered <strong>with</strong> a window <strong>with</strong> l s = 30,<br />

l t = 20. This is to check the statistical isotropy of the WMAP in the modest 20 < l < 40<br />

in the multipole space where certain anomalies have been reported. ILC <strong>with</strong> a 10-<br />

degree-cut (top) has the same BiPS as map C (l s = 30, l t = 20) which explains that the<br />

raising tail of CSSK map is because of the mask.


7.0 119<br />

15<br />

10<br />

5<br />

Wiener<br />

0<br />

-5<br />

15<br />

10<br />

5<br />

CSSK<br />

0<br />

-5<br />

15<br />

10<br />

5<br />

ILC<br />

0<br />

-5<br />

15<br />

10<br />

5<br />

TOH<br />

0<br />

-5<br />

5 10 15 20<br />

Figure 7.4 BiPS measured from 4 different WMAP maps filtered by W S<br />

l (20, 45). From<br />

bottom to up: TOH map, ILC map, CSSK foreground free map <strong>with</strong> a galactic cut,<br />

Wiener filtered map. We see that Wiener filtering has the similar effect as galactic cut.<br />

(But not exactly the same)


7.0 120<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

l=5<br />

l=4<br />

l=3<br />

l=2<br />

l=1<br />

-2 -1 0 1 2<br />

l=5<br />

l=4<br />

l=3<br />

l=2<br />

l=1<br />

-2 -1 0 1 2<br />

Figure 7.5 Probability distribution function for κ 1 to κ 5 constructed from 1000 realizations.<br />

The left panel shows the PDF for the maps filtered <strong>with</strong> W S<br />

l (20, 30) (left panel)<br />

and Wl<br />

G (l s = 40) (right panel). The latter is more skewed, which explains the ∼ 1σ<br />

shift in κ l values shown in Fig. (7.2). The green, magenta and red (circular, pentagonal<br />

and rectangular) points represent ILC, CSSK and TOH maps, respectively. The smooth<br />

solid curves are Gaussian approximations.


7.0 121<br />

0.7<br />

0.6<br />

ILC, W G<br />

TOH, W G<br />

ILC, W S<br />

TOH, W S<br />

0.5<br />

0.4<br />

P(κ l<br />

)<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

0 2 4 6 8 10 12 14 16 18 20<br />

l<br />

Figure 7.6 The probability of the two WMAP based <strong>CMB</strong> maps being SI when filtered<br />

by W S<br />

l (20, 30) and a Gaussian filter W G<br />

l (40).<br />

200<br />

200<br />

100<br />

100<br />

0<br />

0<br />

-100<br />

ILC<br />

-100<br />

ILC<br />

-200<br />

-200<br />

2 4 6 8 10<br />

2 4 6 8 10<br />

200<br />

200<br />

100<br />

100<br />

0<br />

0<br />

-100<br />

Tegmark<br />

-100<br />

Tegmark<br />

-200<br />

-200<br />

2 4 6 8 10<br />

2 4 6 8 10<br />

Figure 7.7 Comparing the measured values of κ l for maps A and B filtered to retain<br />

power only on the lowest multipoles, l = 2 and l = 3 assuming the WMAP theoretical<br />

spectrum WMAPbf (left) and a model spectrum that matches the suppressed power<br />

at the lowest multipoles Shafieloo & Souradeep (2004). The non zero κ l ‘detections’<br />

assuming the WMAP theoretical spectrum become consistent <strong>with</strong> zero for a Cl<br />

T that<br />

has power suppressed at low multipoles.


7.0 122<br />

0.7<br />

ILC<br />

TOH<br />

0.7<br />

ILC<br />

TOH<br />

0.6<br />

0.6<br />

0.5<br />

0.5<br />

0.4<br />

0.4<br />

P(κ l<br />

)<br />

P(κ l<br />

)<br />

0.3<br />

0.3<br />

0.2<br />

0.2<br />

0.1<br />

0.1<br />

0<br />

0 2 4 6 8 10 12 14 16 18 20<br />

l<br />

0<br />

0 2 4 6 8 10 12 14 16 18 20<br />

l<br />

Figure 7.8 probability of the two WMAP based <strong>CMB</strong> maps being SI when filtered<br />

by Wl<br />

G (4) to only keep the lowest multipoles when assuming the WMAP theoretical<br />

spectrum, C T l<br />

lowest multipoles (right).<br />

(left) and a model spectrum that matches the suppressed power at the


7.0 123<br />

100<br />

0<br />

Spergel<br />

50<br />

Spergel<br />

-100<br />

0<br />

-200<br />

0 5 10 15 20<br />

0 5 10 15 20<br />

100<br />

0<br />

ILC<br />

50<br />

ILC<br />

-100<br />

0<br />

-200<br />

0 5 10 15 20<br />

0 5 10 15 20<br />

100<br />

0<br />

Tegmark<br />

50<br />

Tegmark<br />

-100<br />

0<br />

-200<br />

0 5 10 15 20<br />

0 5 10 15 20<br />

20<br />

Spergel<br />

10<br />

Spergel<br />

0<br />

0<br />

0 5 10 15 20<br />

0 5 10 15 20<br />

20<br />

ILC<br />

10<br />

ILC<br />

0<br />

0<br />

0 5 10 15 20<br />

0 5 10 15 20<br />

20<br />

Tegmark<br />

10<br />

Tegmark<br />

0<br />

0<br />

0 5 10 15 20<br />

0 5 10 15 20<br />

Figure 7.9 Measured BiPS for maps A, B and C filtered by bandpass filters W S<br />

l (2, 13)<br />

(top left panel), W S<br />

l (5, 30) (top right), W S<br />

l (20, 30) (bottom left) and W S<br />

l (20, 45) (bottom<br />

right).


7.0 124<br />

100<br />

0<br />

-100<br />

-200<br />

Spergel<br />

0 5 10 15 20<br />

100<br />

0<br />

-100<br />

-200<br />

Spergel<br />

0 5 10 15 20<br />

100<br />

0<br />

-100<br />

-200<br />

0 5 10 15 20<br />

ILC<br />

100<br />

0<br />

-100<br />

-200<br />

ILC<br />

0 5 10 15 20<br />

100<br />

0<br />

-100<br />

-200<br />

Tegmark<br />

0 5 10 15 20<br />

100<br />

0<br />

-100<br />

-200<br />

Tegmark<br />

0 5 10 15 20<br />

0<br />

Spergel<br />

-200<br />

0 5 10 15 20<br />

0<br />

ILC<br />

-200<br />

0 5 10 15 20<br />

0<br />

Tegmark<br />

-200<br />

0 5 10 15 20<br />

Figure 7.10 Measured BiPS for maps A, B and C filtered by Gaussian filters W G<br />

l (40)<br />

(top left panel), W G<br />

l<br />

(18) (top right) and W G (4) (bottom).<br />

l


7.0 125<br />

1<br />

l=5+5<br />

1<br />

l=15+5<br />

0<br />

-4 -2 0 2 4<br />

0<br />

-4 -2 0 2 4<br />

1<br />

l=5+4<br />

1<br />

l=15+4<br />

0<br />

-4 -2 0 2 4<br />

0<br />

-4 -2 0 2 4<br />

1<br />

l=5+3<br />

1<br />

l=15+3<br />

0<br />

-4 -2 0 2 4<br />

0<br />

-4 -2 0 2 4<br />

1<br />

l=5+2<br />

1<br />

l=15+2<br />

0<br />

-4 -2 0 2 4<br />

0<br />

-4 -2 0 2 4<br />

1<br />

l=5+1<br />

1<br />

l=15+1<br />

0<br />

-4 -2 0 2 4<br />

0<br />

-4 -2 0 2 4<br />

1<br />

l=5<br />

1<br />

l=10+5<br />

0<br />

-4 -2 0 2 4<br />

0<br />

-4 -2 0 2 4<br />

1<br />

l=4<br />

1<br />

l=10+4<br />

0<br />

-4 -2 0 2 4<br />

0<br />

-4 -2 0 2 4<br />

1<br />

l=3<br />

1<br />

l=10+3<br />

0<br />

-4 -2 0 2 4<br />

0<br />

-4 -2 0 2 4<br />

1<br />

l=2<br />

1<br />

l=10+2<br />

0<br />

-4 -2 0 2 4<br />

0<br />

-4 -2 0 2 4<br />

1<br />

l=1<br />

1<br />

l=10+1<br />

0<br />

-4 -2 0 2 4<br />

0<br />

-4 -2 0 2 4<br />

Figure 7.11 PDF of κ l for a Gaussian filter Wl<br />

G (40) for 1000 realizations of a SI <strong>CMB</strong><br />

map and the WMAP results. Red point corresponds to the TOH map, green point to<br />

ILC and magenta to the CSSK map.


7.0 126<br />

1<br />

l=5+5<br />

1<br />

l=15+5<br />

0<br />

-4 -2 0 2 4<br />

0<br />

-4 -2 0 2 4<br />

1<br />

l=5+4<br />

1<br />

l=15+4<br />

0<br />

-4 -2 0 2 4<br />

0<br />

-4 -2 0 2 4<br />

1<br />

l=5+3<br />

1<br />

l=15+3<br />

0<br />

-4 -2 0 2 4<br />

0<br />

-4 -2 0 2 4<br />

1<br />

l=5+2<br />

1<br />

l=15+2<br />

0<br />

-4 -2 0 2 4<br />

0<br />

-4 -2 0 2 4<br />

1<br />

l=5+1<br />

1<br />

l=15+1<br />

0<br />

-4 -2 0 2 4<br />

0<br />

-4 -2 0 2 4<br />

1<br />

l=5<br />

1<br />

l=10+5<br />

0<br />

-4 -2 0 2 4<br />

0<br />

-4 -2 0 2 4<br />

1<br />

l=4<br />

1<br />

l=10+4<br />

0<br />

-4 -2 0 2 4<br />

0<br />

-4 -2 0 2 4<br />

1<br />

l=3<br />

1<br />

l=10+3<br />

0<br />

-4 -2 0 2 4<br />

0<br />

-4 -2 0 2 4<br />

1<br />

l=2<br />

1<br />

l=10+2<br />

0<br />

-4 -2 0 2 4<br />

0<br />

-4 -2 0 2 4<br />

1<br />

l=1<br />

1<br />

l=10+1<br />

0<br />

-4 -2 0 2 4<br />

0<br />

-4 -2 0 2 4<br />

Figure 7.12 PDF of κ l for a band pass filter W S<br />

l (20, 30) for 1000 realizations of a SI <strong>CMB</strong><br />

map and the WMAP results. Red point corresponds to the TOH map, green point to<br />

ILC and magenta to the CSSK map.


Chapter 8<br />

Summary and future directions<br />

We have studied statistics of the <strong>CMB</strong> anisotropy as a Gaussian random field on a<br />

sphere. We studied statistical isotropy (SI) of <strong>CMB</strong> temperature anisotropy and used<br />

that to draw interesting conclusions on the physics, topology and large scale structure<br />

of the universe as well as properties of the systematics in <strong>CMB</strong> experiments. Statistical<br />

isotropy is usually assumed in almost all <strong>CMB</strong> studies. But there are certain mechanisms<br />

that violate this condition. This thesis is dedicated to the study of this fundamental<br />

assumption in <strong>CMB</strong>. We have formulated the problem mathematically, constructed a<br />

quantitative method to describe SI violation, studied possible ways of violation of SI<br />

condition and finally examined this fundamental assumption using the most recent <strong>CMB</strong><br />

anisotropy data.<br />

The SI of the <strong>CMB</strong> anisotropy has been under scrutiny after the release of the first<br />

year of WMAP data. We use the BiPS which is sensitive to structures and patterns<br />

in the underlying two-point correlation function as a statistical tool of searching for<br />

departures from SI. We carry out a BiPS analysis of WMAP full sky maps. We find no<br />

strong evidence for SI violation in the WMAP <strong>CMB</strong> anisotropy maps considered here.<br />

We have verified that our null results are consistent <strong>with</strong> measurements on simulated SI<br />

maps. The BiPS measurement reported here is a Bayesian estimate of the conditional<br />

probability of SI (for each κ l of the BiPS) given an underlying theoretical spectrum<br />

Cl T . We point out that the excess power in the C l<br />

T <strong>with</strong> respect to the measured C l<br />

from WMAP at the lowest multipoles tends to indicate mild deviations from SI. BiPS<br />

measurements are shown to be consistent <strong>with</strong> SI assuming an alternate model Cl<br />

T that<br />

127


8.0 128<br />

is consistent <strong>with</strong> suppressed power on low multipoles. The full sky maps and the<br />

restriction to low l < 60 (where instrumental noise is sub-dominant) permits the use of<br />

our analytical bias subtraction and error estimates.<br />

There are strong theoretical motivations for hunting for SI violation in the <strong>CMB</strong><br />

anisotropy. The possibility of compact spaces in cosmic topology is a theoretically well<br />

motivated possibility that has also been observationally targeted Ellis (1971); Lachieze-<br />

Rey & Luminet (1995); Levin (2002); Linde (2004). The breakdown of statistical<br />

homogeneity and isotropy of cosmic perturbations is a generic feature of ultra large scale<br />

structure of the cosmos, in particular, of non trivial cosmic topology Bond, Pogosyan &<br />

Souradeep (1998, 2000). The underlying correlation patterns in the <strong>CMB</strong> anisotropy in<br />

a multiply connected universe is related to the symmetry of the Dirichlet domain. The<br />

BiPS expected in flat, toroidal models of the universe has been computed and shown<br />

to be related to the principle directions in the Dirichlet domain Hajian & Souradeep<br />

(2003a). As a tool for constraining cosmic topology, the BiPS has the advantage of being<br />

independent of the overall orientation of the Dirichlet domain <strong>with</strong> respect to the sky.<br />

Hence, the null result of BiPS can have important implication for cosmic topology. This<br />

approach complements direct search for signature of cosmic topology Cornish, Spergel<br />

& Starkman (1998); de Oliveira-Costa et al. (1996) and our results are consistent<br />

<strong>with</strong> the absence of the matched circles and the null S-map test of the WMAP <strong>CMB</strong><br />

maps Cornish et al. (2003); de Oliveira-Costa et al. (2003). Full Bayesian likelihood<br />

comparison to the data of specific cosmic topology models is another approach that has<br />

applied to COBE-DMR data Bond, Pogosyan & Souradeep (1998, 2000). We also study<br />

Poincare Dodecahedral spaces suggested as a possible explanation of low quadrupole of<br />

the WMAP data by Luminet et al. (2003). We show that the Poincaré Dodecahedron<br />

<strong>with</strong> Ω k = 0.013 breaks down the statistical isotropy at an observable level which was<br />

not observed in the WMAP data. Our results are consistent <strong>with</strong> the complete Bayesian<br />

likelihood analysis of Poincaré Dodecahedral spaces [Hajian et al. (2006)].<br />

Discovery of a hidden pattern in the WMAP data was recently reported by Jaffe et<br />

al. (2005) which according to the authors was due to a rotating universe template (a<br />

Bianchi pattern). The model studied by Jaffe et al. (2005) was examined in this thesis.<br />

We show that hiding a Bianchi template in a ΛCDM <strong>CMB</strong> anisotropy temperature map,


8.0 129<br />

will lead to measurable non-zero bipolar power spectra. The null BiPS of WMAP can<br />

be used to put tight constraints on the existence of such hidden patterns in the WMAP<br />

data.<br />

The null BiPS results also has implications for the observation and data analysis<br />

techniques used to create the <strong>CMB</strong> anisotropy maps. Observational artifacts such as<br />

non-circular beam, inhomogeneous noise correlation, residual striping patterns, etc. are<br />

potential sources of SI breakdown. Our null BiPS results confirm that these artifacts<br />

do not significantly contribute to the maps studied here. Foreground residuals can also<br />

be sources of SI breakdown. The extent to which BiPS probes foreground residuals is<br />

yet to be fully studied and explored. We do not see any significant effect of the residual<br />

foregrounds in ILC and the TOH maps as it was mentioned by Eriksen et al. (2004c).<br />

This can not be necessarily called a discrepancy between the two results unless we have<br />

a model that can predict what should have been seen in the BiPS. However the question<br />

which remains is if the signal is strong enough and whether the effect smeared out in<br />

bipolar multipole space <strong>with</strong>in our angular l-space window. On the other hand, we have<br />

verified that BiPS does show a strong signal for the residuals from foregrounds in the<br />

Wiener filtered map.<br />

In the very near future we are going to have <strong>CMB</strong> polarization sky maps. The wealth<br />

of information in the <strong>CMB</strong> polarization field will enable us to test and characterize the<br />

initial perturbations and inflationary mechanisms <strong>with</strong> great precision. The cosmological<br />

polarized microwave radiation in a simply connected universe is expected to be<br />

statistically isotropic. This is a very important feature which allows us to fully describe<br />

the field by its power spectrum. Violation of SI in <strong>CMB</strong> polarization maps is an interesting<br />

issue which is going to be very important soon. Importance of having statistical<br />

tests of departures from SI for <strong>CMB</strong> polarization maps lies not only in the theoretical<br />

motivations discussed above but also in testing the cleaned <strong>CMB</strong> polarization maps for<br />

residuals from polarized foreground emission. Unlike the foregrounds in temperature<br />

anisotropies, polarized foreground emissions which are mostly from our galaxy and also<br />

from extra-galactic sources on large scales (where we may see the primordial B-mode<br />

due to inflationary gravitational waves) is likely to dominate the signal at all frequencies.<br />

A strong tool to distinguish between the primordial polarized radiation and polarized


8.0 130<br />

foreground emissions is the test of SI. BiPS, by construction is applicable to polarized<br />

maps. We can check for departures from SI in E and B polarization modes as well as<br />

the cross terms such as TE by computing BiPS for them.


AppendixA<br />

Detailed Steps of Calculations<br />

Gaussianity of <strong>CMB</strong> anisotropy<br />

The temperature anisotropy at a given space-time point (x, τ) can be written as a<br />

superposition of plane waves in k-space<br />

∫<br />

∆T (x, ˆn, τ) = d 3 k e ik·x ∆T (k, ˆn, τ).<br />

(A.1)<br />

As discussed in Ma & Bertschinger (1995), the dependence on ˆn arises only through<br />

k · ˆn. Therefore ∆T (k, ˆn, τ) may be represented as<br />

∞∑<br />

∆T (k, ˆn, τ) = (−i) l (2l + 1)∆T l (k, τ)P l (k · ˆn).<br />

(A.2)<br />

l=0<br />

The anisotropy coefficients ∆T l (k, τ) are random variables whose amplitudes and phases<br />

depend on the initial perturbations and can be written as<br />

∆T l (k, τ) = φ 0 (k)∆T l (k, τ),<br />

(A.3)<br />

where φ 0 (k) is the initial potential perturbation and ∆T l (k, τ) is the photon transfer<br />

function, i.e. solution of Boltzmann equation <strong>with</strong> φ 0 (k) = 1. In general, ∆T l (k, τ) is<br />

given by the integral solution<br />

∆T l (β) =<br />

∫ τ0<br />

0<br />

dτΦ l β(τ 0 − τ)S(β, τ),<br />

(A.4)<br />

in which S(β, τ) is the source function, Φ l β is the ultra-spherical Bessel function and<br />

β = k − K, where K = H0(Ω 2 0 − 1) is the curvature Zaldarriaga et al. (1998). In the<br />

131


A.0 132<br />

flat case, the above expression has a very simple form<br />

∆T (k) =<br />

∫ τ0<br />

0<br />

dτ e i k·ˆn<br />

k (τ−τ 0) S(k, τ), (A.5)<br />

and hence, the temperature anisotropies are given by<br />

∫<br />

∫ τ0<br />

∆T (x, ˆn) = d 3 k e ik·x φ 0 (k)[ dτ e i k·ˆn<br />

k (τ−τ0) S(k, τ)]. (A.6)<br />

Equation (A.6) clearly shows that Gaussian perturbations in primordial gravitational<br />

potential result a Gaussian <strong>CMB</strong> temperature anisotropy field.<br />

0<br />

SI condition in harmonic space<br />

It was shown in eqn.<br />

(2.17) that statistical isotropy as invariance of two point<br />

correlation under rotation, translates to a diagonal covariance matrix in harmonic space.<br />

To prove this we substitute the spherical harmonic coefficients, a lm , from eqn. (2.8) in<br />

the covariance matrix to express it in terms of the two point correlation<br />

∫ ∫<br />

〈a lm a ∗ l ′ m ′〉 = dΩˆn dΩˆn ′Ylm ∗ (ˆn)Y l ′ m ′(ˆn′ )〈∆T (ˆn)∆T (ˆn ′ )〉<br />

∫ ∫<br />

= dΩˆn dΩˆn ′Ylm(ˆn)Y ∗<br />

l ′ m ′(ˆn′ )C(ˆn, ˆn ′ ).<br />

(A.7)<br />

Under the condition of SI, eqn. (2.13), and after making use of the spherical harminc<br />

expansion of Legendre polynomials, eqn. (G.6), in eqn. (2.14), the above expression<br />

will become<br />

∫ ∫<br />

〈a lm a ∗ l ′ m ′〉 = dΩˆn dΩˆn ′Ylm(ˆn)Y ∗<br />

lm (ˆn ′ ) ∑ l 1<br />

= ∑ ∑<br />

C l1<br />

l 1<br />

m 1<br />

δ ll1 δ l ′ l 1<br />

δ mm1 δ m ′ m 1<br />

= C l δ ll ′δ mm ′,<br />

which is the SI condition in harmonic space, eqn.<br />

C l1<br />

∑<br />

m 1<br />

Y l1 m 1<br />

(ˆn)Y l1 m 1<br />

(ˆn ′ ) (A.8)<br />

(2.17), and in the last line, the<br />

orthonormality relation of spherical harmonics, eqn. (G.2) has been used.


A.0 133<br />

Cosmic Variance Surplus<br />

The angular power spectrum, C l , is an “isotropized” measure. The isolation of SI<br />

implies that the C l do not contain the entire information contents of the anisotropy<br />

field. This leads to the surplus of cosmic variance of C l in cases where SI is violated.<br />

Here is a proof of this. The estimator of angular power spectrum is given by<br />

˜C l = 1<br />

2l + 1<br />

l∑<br />

m=−l<br />

a lm a ∗ lm,<br />

(A.9)<br />

for a Gaussian isotropic theory, this has the cosmic variance<br />

σ 2 SI( ˜C l ) =<br />

2C2 l<br />

2l + 1 .<br />

For a general theory, the cosmic variance of the above estimator is given by<br />

σ 2 ( ˜C l ) =<br />

2<br />

(2l + 1) [∑ ∑<br />

〈a 2 lm a ∗ lm ′〉〈a lma ∗ lm ′〉∗ δ mm ′<br />

m m ′<br />

+ ∑ ∑<br />

〈a lm a ∗ lm ′〉〈a lma ∗ lm ′〉∗ ]<br />

m≠m ′ m ′<br />

= 2C2 l<br />

2l + 1 + 2<br />

(2l + 1) 2 ∑<br />

m≠m ′ ∑<br />

= σ 2 SI ( ˜C l ) + σ 2 SA .<br />

m ′ 〈a lm a ∗ lm ′〉〈a lma ∗ lm ′〉∗<br />

(A.10)<br />

(A.11)<br />

The second term (σSA 2 ) is the sum of non-negative numbers and hence is non-negative.<br />

Therefore the cosmic variance is always greater than or equal to that of SI models<br />

σ 2 ( ˜C l ) ≥ σ 2 SI ( ˜C l ).<br />

(A.12)<br />

The fact that cosmic variance of angular power spectrum is minimized for SI theories,<br />

was shown in Ferreira & Magueijo (1997). Also Bond, Pogosyan & Souradeep<br />

(2000) in a detailed calculations for statistically anisotropic theories showed that the incompleteness<br />

of the information contained in the C l is reflected in the enhanced cosmic<br />

variance. This can be seen by splitting the two point correlation into an isotropic part<br />

determined by C l and an anisotropic term containing the remainder<br />

C(ˆn 1 , ˆn 2 ) = C SI (ˆn 1 , ˆn 2 ) + C SA (ˆn 1 , ˆn 2 ).<br />

(A.13)


A.0 134<br />

It is then straightfoward to calculate the cosmic variace due to each part and the final<br />

expression as it is given in Bond, Pogosyan & Souradeep (2000) is<br />

σ 2 ( ˜C l ) =<br />

2C2 l<br />

2l + 1 + l2 (l + 1) 2 ∫<br />

32π 4<br />

dΩˆn1<br />

∫<br />

∫<br />

dΩˆn2 [<br />

dΩˆn3 C SA (ˆn 1 , ˆn 2 )P l (ˆn 1 · ˆn 2 )] 2 .<br />

(A.14)<br />

In the above expression, the second term is a non-negative term and therefore we can<br />

again see that in isotropic theories, the cosmic variance of the angular power spectrum<br />

has its smallest value.<br />

A lM<br />

l 1 l 2<br />

are linear combinations of 〈a lm a ∗ l ′ m ′ 〉<br />

The BiPoSH coefficients, A lM<br />

l 1 l 2<br />

, are linear combinations a lm .<br />

To see that, in the<br />

definition of two point correlation we can substitute for ∆T/T from eqn. (2.7), which<br />

will lead to the expansion of C(ˆn 1 , ˆn 2 ) in terms of spherical harmonics<br />

C(ˆn 1 , ˆn 2 ) = 〈∆T (ˆn 1 )∆T (ˆn 2 )〉 (A.15)<br />

= ∑ ∑<br />

〈a lm a ∗ l ′ m ′〉Y lm(ˆn 1 )Yl ∗ ′ m ′(ˆn 2).<br />

lm l ′ m ′<br />

This can be substituted in the definition of A lM<br />

l 1 l 2<br />

and eqn. (2.20) can be written as<br />

∫ ∫<br />

A lM<br />

l 1 l 2<br />

= dΩ 1 dΩ 2 C(ˆn 1 , ˆn 2 ){Y l1 (ˆn 1 ) × Y l2 (ˆn 2 )} ∗ lM<br />

(A.16)<br />

∫ ∫<br />

∑ ∑<br />

= dΩ 1 dΩ 2 〈a lm a ∗ l ′ m ′〉Y lm(ˆn 1 )Yl ∗ ′ m ′(ˆn 2) ∑ Cl lM<br />

1 m 1 l 2 m 2<br />

Yl ∗<br />

1 m 1<br />

(ˆn 1 )Yl ∗<br />

2 m 2<br />

(ˆn 2 )<br />

lm l ′ m ′ m 1 m<br />

∫ ∫<br />

2<br />

∑ ∑<br />

= dΩ 1 dΩ 2 〈a lm a ∗ l ′ m ′〉Y lm(ˆn 1 )Yl ∗ ′ m ′(ˆn 2)<br />

lm l ′ m ′<br />

×<br />

∑<br />

m 1 m 2<br />

C lM<br />

l 1 m 1 l 2 m 2<br />

Y ∗<br />

l 1 m 1<br />

(ˆn 1 )(−1) m 2<br />

Y l2 −m 2<br />

(ˆn 2 )<br />

= ∑ m 1 m 2<br />

(−1) m 2<br />

〈a l1 m 1<br />

a ∗ l 2 −m 2<br />

〉C lM<br />

l 1 m 1 l 2 m 2<br />

By changing the dummy variable m 2 → −m 2 we will get<br />

A lM<br />

l 1 l 2<br />

= ∑ m 1 m 2<br />

〈a l1 m 1<br />

a ∗ l 2 m 2<br />

〉(−1) m 2<br />

C lM<br />

l 1 m 1 l 2 −m 2<br />

.<br />

(A.17)


A.0 135<br />

SI condition in bipolar harmonic space<br />

If we substitute the covariance matrix of a SI theory, eqn. (2.17), into the expression<br />

for A lM<br />

ll , eqn. (A.17), we would get<br />

′<br />

A lM<br />

ll ′ = ∑ mm ′ C l δ ll ′δ m−m ′(−1) m′ C lM<br />

lml ′ m ′<br />

∑<br />

= C l δ ll ′δ M0 (−1) m C lM<br />

m<br />

lml−m.<br />

(A.18)<br />

Using summation rules of Clebsch-Gordan equations, eqn. (G.11), the above expression<br />

will become<br />

A lM<br />

ll ′ = (−1)l C l (2l + 1) 1/2 δ ll ′ δ l0 δ M0 . (A.19)<br />

This is the SI condition in bipolar harmonic space and it arises as we carry out the sum<br />

over m.<br />

Simplifying the real-space expression for BiPS<br />

The BiPS in real-space is given by the following expression.<br />

κ l = ( 2l + 1 ∫ ∫ ∫<br />

∫<br />

8π 2 )2 dΩ 1 dΩ 2 dR 1 χ l (R 1 ) dR 2 χ l (R 2 )C(R 1ˆn 1 , R 1ˆn 2 )C(R 2ˆn 1 , R 2ˆn 2 )<br />

(A.20)<br />

Now if we change the variables in the following way<br />

R 1ˆn 1 = ˆn 1<br />

(A.21)<br />

R 1ˆn 2 = ˆn 2<br />

We will have<br />

κ l = ( 2l + 1 ∫<br />

) 2<br />

8π 2<br />

dΩ 1<br />

∫<br />

dΩ 2<br />

∫<br />

∫<br />

dR 1 χ l (R 1 )<br />

dR 2 χ l (R 2 )C(ˆn 1 , ˆn 2 )C(R 2 R −1<br />

1 ˆn 1, R 2 R −1<br />

1 ˆn 2),<br />

(A.22)<br />

and <strong>with</strong><br />

R = R 2 R −1<br />

1 (A.23)<br />

we will get<br />

κ l = ( 2l + 1 ∫<br />

) 2<br />

8π 2<br />

dΩ 1<br />

∫<br />

dΩ 2<br />

∫<br />

∫<br />

dR 1 χ l (R 1 )<br />

dRχ l (RR 1 )C(ˆn 1 , ˆn 2 )C(Rˆn 1 , Rˆn 2 ).<br />

(A.24)


A.0 136<br />

Using eqn. (G.9) we can write<br />

∫<br />

dR 1 χ l (R 1 )χ l (RR 1 ) = ∑ m 1<br />

∑<br />

mm ′ ∫<br />

= ∑ m 1<br />

∑<br />

=<br />

=<br />

8π 2<br />

2l + 1<br />

mm ′ D l mm<br />

dR 1 D ∗l<br />

m 1 m 1<br />

(R 1 )D l mm ′(R)Dl m ′ m(R 1 ) (A.25)<br />

8π2<br />

′(R)<br />

2l + 1 δ m 1 m ′δ m 1 m<br />

∑<br />

Dm l 1 m 1<br />

(R)<br />

m 1<br />

8π 2<br />

2l + 1 χ l(R).<br />

So, the κ l can be written in this form:<br />

κ l = 2l + 1 ∫ ∫<br />

∫<br />

dΩ<br />

8π 2 1 dΩ 2 C(ˆn 1 , ˆn 2 )<br />

dRχ l (R)C(Rˆn 1 , Rˆn 2 ).<br />

(A.26)<br />

Computing the Cosmic Variance of BiPS<br />

In section 2.6 we presented the analytical expression for the cosmic variance of BiPS.<br />

Here we show the details of our calculation. The cosmic variance, 〈˜κ 2 l 〉 − 〈˜κ l〉 2 , has 96<br />

terms similar to the one shown in eqn. (2.44) but <strong>with</strong> different combinations of R’s and<br />

ˆn’s. Among them, 40 terms contain at least one term like C(Rˆn 1 , Rˆn 2 ). These terms<br />

only contribute to κ 0 because when SI holds, C(Rˆn 1 , Rˆn 2 ) = C(ˆn 1 , ˆn 2 ) and hence the<br />

integral in eqn. (A.26) will be proportional to δ l0 , which is not in the interest of us.<br />

So, we can ignore them and we will be left <strong>with</strong> 56 terms (more details on calculation<br />

of these terms are given in Appendix B). These terms, depending on the number of<br />

connected correlations in them, can be classified into 4 major groups:<br />

1. One-loop terms, such as<br />

C(ˆn 1 , R ′ˆn 3 )C(ˆn 3 , Rˆn 2 )C(ˆn 2 , R ′ˆn 4 )C(ˆn 4 , Rˆn 1 ).<br />

(A.27)<br />

These terms get reduced to expressions <strong>with</strong> a C 4 l<br />

coefficients (summation symbols are omitted for brevity),<br />

factor and 4 Clebsch-Gordan<br />

C 4 l 1<br />

C lM<br />

l 1 −m 1 l 1 −m 3<br />

C lM<br />

l 1 m 5 l 1 m 7<br />

C lM ′<br />

l 1 m 7 l 1 m 5<br />

C lM ′<br />

l 1 −m 1 l 1 −m 3<br />

.<br />

(A.28)<br />

There are 24 terms like this but <strong>with</strong> different combinations of indices.


A.0 137<br />

2. Two-loop terms – type I, such as<br />

C(ˆn 1 , ˆn 3 )C(R ′ˆn 3 , Rˆn 2 )C(ˆn 2 , Rˆn 1 )C(ˆn 4 , R ′ ˆn 4 ).<br />

(A.29)<br />

These terms get reduced to expressions <strong>with</strong> a C 3 l 1<br />

C l2 factor and 4 Clebsch-Gordan<br />

coefficients<br />

C 3 l 1<br />

C l7 C lM<br />

l 1 −m 1 l 1 −m 3<br />

C lM<br />

l 1 −m 3 l 1 m 5<br />

C lM ′<br />

l 1 m 1 l 7 −m 7<br />

C lM ′<br />

l 1 −m 5 l 7 −m 7<br />

.<br />

There are 16 terms like this but <strong>with</strong> different combinations of indices.<br />

3. Two-loop terms – type II, such as<br />

(A.30)<br />

C(ˆn 1 , R ′ˆn 4 )C(ˆn 4 , Rˆn 1 )C(ˆn 2 , R ′ˆn 3 )C(Rˆn 2 , ˆn 3 ).<br />

(A.31)<br />

These terms get reduced to expressions <strong>with</strong> a C 2 l 1<br />

C 2 l 2<br />

factor and 4 Clebsch-Gordan<br />

coefficients<br />

C 2 l 1<br />

C 2 l 3<br />

C lM<br />

l 1 −m 1 l 3 −m 3<br />

C lM<br />

l 1 m 5 l 3 m 7<br />

C lM ′<br />

l 1 m 5 l 3 m 7<br />

C lM ′<br />

l 1 −m 1 l 3 −m 3<br />

.<br />

There are 8 terms like this but <strong>with</strong> different combinations of indices.<br />

4. Three-loop terms, such as<br />

(A.32)<br />

C(ˆn 1 , ˆn 3 )C(R ′ˆn 3 , Rˆn 1 )C(ˆn 2 , Rˆn 2 )C(ˆn 4 , R ′ˆn 4 ).<br />

(A.33)<br />

These terms get reduced to expressions <strong>with</strong> a C 2 l 1<br />

C l2 C l3<br />

Gordan coefficients<br />

factor and 4 Clebsch-<br />

C 2 l 1<br />

C l3 C l7 C lM<br />

l 3 −m 3 l 1 −m 1<br />

C lM<br />

l 3 −m 3 l 1 m 5<br />

C lM ′<br />

l 1 m 1 l 7 −m 7<br />

C lM ′<br />

l 1 −m 5 l 7 −m 7<br />

.<br />

(A.34)<br />

There are 8 terms like this but <strong>with</strong> different combinations of indices.<br />

Using symmetry properties of Clebsch-Gordan coefficients, eqn. (G.10), and their<br />

summation rules, eqn. (G.11), it is possible to simplify these 56 terms immensely. After<br />

tedious algebra we would obtain the final expression for the kernel of cosmic variance of<br />

the bipolar power spectrum:<br />

4 Cl 4 (2l + 1)2<br />

1<br />

[2<br />

2l 1 + 1<br />

+ (−1)l (2l + 1) + (1 + 2(−1) l )F l<br />

ll ] {l 1l 1 l} +<br />

4 C 2 l 1<br />

C 2 l 2<br />

[(2l + 1) + F l<br />

l 1 l 2<br />

] {l 1 l 2 l} +<br />

8 C 2 l 1<br />

C l2 C l3<br />

(2l + 1) 2<br />

2l 1 + 1 {l 1l 2 l} {l 1 l 3 l} +<br />

16 (−1) l Cl 3 (2l + 1) 2<br />

1<br />

C l3<br />

2l 1 + 1 {l 1l 1 l} {l 1 l 3 l}<br />

(A.35)


A.0 138<br />

where the 3j-symbol {l 1 l 2 l} is given by<br />

⎧<br />

⎨ 1 if l 1 + l 2 + l is integer and |l 1 − l 2 | ≤ l ≤ (l 1 + l 2 ),<br />

{l 1 l 2 l} =<br />

⎩ 0 otherwise,<br />

and F l<br />

l 1 l 3<br />

is defined as<br />

F l<br />

l 1 l 3<br />

=<br />

l 1<br />

∑<br />

m 1 m 2 =−l 1<br />

l 3<br />

∑<br />

l∑<br />

m 3 m 4 =−l 3 M,M ′ =−l<br />

C lM<br />

l 1 −m 1 l 3 −m 3<br />

C lM<br />

l 1 m 2 l 3 m 4<br />

C lM ′<br />

l 3 m 4 l 1 m 1<br />

C lM ′<br />

l 3 −m 3 l 1 −m 2<br />

.<br />

(A.36)<br />

This should be summed over all indices except l to result the analytical expression for<br />

the cosmic variance given in eqn. (2.45).


AppendixB<br />

Cosmic Variance of BiPS, Handling<br />

Gaussian 8 Point Functions<br />

As it is shown in eqn. (2.42), to compute the cosmic variance (CV) of BiPS, we<br />

must deal <strong>with</strong> an eight point correlation. Assuming Gaussiantiy of the field we can<br />

rewrite the eight point correlation in terms of two point correlations. This will give us<br />

(8 − 1)!! = 7 × 5 × 3 = 105 terms. These 105 terms are given below:<br />

x1x2 x3x4 x5x6 x7x8 + x1x2 x3x4 x5x7 x6x8 + x1x2 x3x4 x5x8 x6x7 +<br />

x1x2 x3x5 x4x6 x7x8 + x1x2 x3x5 x4x7 x6x8 + x1x2 x3x5 x4x8 x6x7 +<br />

x1x2 x3x6 x4x5 x7x8 + x1x2 x3x6 x4x7 x5x8 + x1x2 x3x6 x4x8 x5x7 +<br />

x1x2 x3x7 x4x5 x6x8 + x1x2 x3x7 x4x6 x5x8 + x1x2 x3x7 x4x8 x5x6 +<br />

x1x2 x3x8 x4x5 x6x7 + x1x2 x3x8 x4x6 x5x7 + x1x2 x3x8 x4x7 x5x6 +<br />

x1x3 x2x4 x5x6 x7x8 + x1x3 x2x4 x5x7 x6x8 + x1x3 x2x4 x5x8 x6x7 +<br />

x1x3 x2x5 x4x6 x7x8 + x1x3 x2x5 x4x7 x6x8 + x1x3 x2x5 x4x8 x6x7 +<br />

x1x3 x2x6 x4x5 x7x8 + x1x3 x2x6 x4x7 x5x8 + x1x3 x2x6 x4x8 x5x7 +<br />

x1x3 x2x7 x4x5 x6x8 + x1x3 x2x7 x4x6 x5x8 + x1x3 x2x7 x4x8 x5x6 +<br />

x1x3 x2x8 x4x5 x6x7 + x1x3 x2x8 x4x6 x5x7 + x1x3 x2x8 x4x7 x5x6 +<br />

x1x4 x2x3 x5x6 x7x8 + x1x4 x2x3 x5x7 x6x8 + x1x4 x2x3 x5x8 x6x7 +<br />

x1x4 x2x5 x3x6 x7x8 + x1x4 x2x5 x3x7 x6x8 + x1x4 x2x5 x3x8 x6x7 +<br />

x1x4 x2x6 x3x5 x7x8 + x1x4 x2x6 x3x7 x5x8 + x1x4 x2x6 x3x8 x5x7 +<br />

x1x4 x2x7 x3x5 x6x8 + x1x4 x2x7 x3x6 x5x8 + x1x4 x2x7 x3x8 x5x6 +<br />

x1x4 x2x8 x3x5 x6x7 + x1x4 x2x8 x3x6 x5x7 + x1x4 x2x8 x3x7 x5x6 +<br />

x1x5 x2x3 x4x6 x7x8 + x1x5 x2x3 x4x7 x6x8 + x1x5 x2x3 x4x8 x6x7 +<br />

x1x5 x2x4 x3x6 x7x8 + x1x5 x2x4 x3x7 x6x8 + x1x5 x2x4 x3x8 x6x7 +<br />

x1x5 x2x6 x3x4 x7x8 + x1x5 x2x6 x3x7 x4x8 + x1x5 x2x6 x3x8 x4x7 +<br />

x1x5 x2x7 x3x4 x6x8 + x1x5 x2x7 x3x6 x4x8 + x1x5 x2x7 x3x8 x4x6 +<br />

x1x5 x2x8 x3x4 x6x7 + x1x5 x2x8 x3x6 x4x7 + x1x5 x2x8 x3x7 x4x6 +<br />

139


B.0 140<br />

x1x6 x2x3 x4x5 x7x8 + x1x6 x2x3 x4x7 x5x8 + x1x6 x2x3 x4x8 x5x7 +<br />

x1x6 x2x4 x3x5 x7x8 + x1x6 x2x4 x3x7 x5x8 + x1x6 x2x4 x3x8 x5x7 +<br />

x1x6 x2x5 x3x4 x7x8 + x1x6 x2x5 x3x7 x4x8 + x1x6 x2x5 x3x8 x4x7 +<br />

x1x6 x2x7 x3x4 x5x8 + x1x6 x2x7 x3x5 x4x8 + x1x6 x2x7 x3x8 x4x5 +<br />

x1x6 x2x8 x3x4 x5x7 + x1x6 x2x8 x3x5 x4x7 + x1x6 x2x8 x3x7 x4x5 +<br />

x1x7 x2x3 x4x5 x6x8 + x1x7 x2x3 x4x6 x5x8 + x1x7 x2x3 x4x8 x5x6 +<br />

x1x7 x2x4 x3x5 x6x8 + x1x7 x2x4 x3x6 x5x8 + x1x7 x2x4 x3x8 x5x6 +<br />

x1x7 x2x5 x3x4 x6x8 + x1x7 x2x5 x3x6 x4x8 + x1x7 x2x5 x3x8 x4x6 +<br />

x1x7 x2x6 x3x4 x5x8 + x1x7 x2x6 x3x5 x4x8 + x1x7 x2x6 x3x8 x4x5 +<br />

x1x7 x2x8 x3x4 x5x6 + x1x7 x2x8 x3x5 x4x6 + x1x7 x2x8 x3x6 x4x5 +<br />

x1x8 x2x3 x4x5 x6x7 + x1x8 x2x3 x4x6 x5x7 + x1x8 x2x3 x4x7 x5x6 +<br />

x1x8 x2x4 x3x5 x6x7 + x1x8 x2x4 x3x6 x5x7 + x1x8 x2x4 x3x7 x5x6 +<br />

x1x8 x2x5 x3x4 x6x7 + x1x8 x2x5 x3x6 x4x7 + x1x8 x2x5 x3x7 x4x6 +<br />

x1x8 x2x6 x3x4 x5x7 + x1x8 x2x6 x3x5 x4x7 + x1x8 x2x6 x3x7 x4x5 +<br />

x1x8 x2x7 x3x4 x5x6 + x1x8 x2x7 x3x5 x4x6 + x1x8 x2x7 x3x6 x4x5<br />

Where by x1x2 I mean the two point correlation between 1 and 2, i.e. 〈∆T (R 1ˆq 1 )∆T (R 1ˆq 2 )〉.<br />

On the other hand < κ l > has a 4 point correlation in it<br />

< ∆T (R 1ˆq 1 )∆T (R 1ˆq 2 )∆T (R 2ˆq 1 )∆T (R 2ˆq 2 ) > (B.1)<br />

which can be written as, say, x1x2 x3x4 + x1x3 x2x4 + x1x4 x2x3; so < κ l > 2 will be<br />

(x1x2 x3x4 + x1x3 x2x4 + x1x4 x2x3) (x5x6 x7x8 + x5x7 x6x8 + x5x8 x6x7) =<br />

x1x4 x2x3 x5x8 x6x7 + x1x3 x2x4 x5x8 x6x7 + x1x2 x3x4 x5x8 x6x7 +<br />

x1x4 x2x3 x5x7 x6x8 + x1x3 x2x4 x5x7 x6x8 + x1x2 x3x4 x5x7 x6x8 +<br />

x1x4 x2x3 x5x6 x7x8 + x1x3 x2x4 x5x6 x7x8 + x1x2 x3x4 x5x6 x7x8<br />

If we subtract these two, we will get:<br />

< κ 2 l > − < κ l > 2 =<br />

x1x8 x2x7 x3x6 x4x5 + x1x7 x2x8 x3x6 x4x5 + x1x8 x2x6 x3x7 x4x5 +<br />

x1x6 x2x8 x3x7 x4x5 + x1x7 x2x6 x3x8 x4x5 + x1x6 x2x7 x3x8 x4x5 +<br />

x1x8 x2x7 x3x5 x4x6 + x1x7 x2x8 x3x5 x4x6 + x1x8 x2x5 x3x7 x4x6 +<br />

x1x5 x2x8 x3x7 x4x6 + x1x7 x2x5 x3x8 x4x6 + x1x5 x2x7 x3x8 x4x6 +<br />

x1x8 x2x6 x3x5 x4x7 + x1x6 x2x8 x3x5 x4x7 + x1x8 x2x5 x3x6 x4x7 +<br />

x1x5 x2x8 x3x6 x4x7 + x1x6 x2x5 x3x8 x4x7 + x1x5 x2x6 x3x8 x4x7 +<br />

x1x7 x2x6 x3x5 x4x8 + x1x6 x2x7 x3x5 x4x8 + x1x7 x2x5 x3x6 x4x8 +<br />

x1x5 x2x7 x3x6 x4x8 + x1x6 x2x5 x3x7 x4x8 + x1x5 x2x6 x3x7 x4x8 +<br />

x1x8 x2x7 x3x4 x5x6 + x1x7 x2x8 x3x4 x5x6 + x1x8 x2x4 x3x7 x5x6 +


B.0 141<br />

x1x4 x2x8 x3x7 x5x6 + x1x7 x2x4 x3x8 x5x6 + x1x4 x2x7 x3x8 x5x6 +<br />

x1x8 x2x3 x4x7 x5x6 + x1x3 x2x8 x4x7 x5x6 + x1x2 x3x8 x4x7 x5x6 +<br />

x1x7 x2x3 x4x8 x5x6 + x1x3 x2x7 x4x8 x5x6 + x1x2 x3x7 x4x8 x5x6 +<br />

x1x8 x2x6 x3x4 x5x7 + x1x6 x2x8 x3x4 x5x7 + x1x8 x2x4 x3x6 x5x7 +<br />

x1x4 x2x8 x3x6 x5x7 + x1x6 x2x4 x3x8 x5x7 + x1x4 x2x6 x3x8 x5x7 +<br />

x1x8 x2x3 x4x6 x5x7 + x1x3 x2x8 x4x6 x5x7 + x1x2 x3x8 x4x6 x5x7 +<br />

x1x6 x2x3 x4x8 x5x7 + x1x3 x2x6 x4x8 x5x7 + x1x2 x3x6 x4x8 x5x7 +<br />

x1x7 x2x6 x3x4 x5x8 + x1x6 x2x7 x3x4 x5x8 + x1x7 x2x4 x3x6 x5x8 +<br />

x1x4 x2x7 x3x6 x5x8 + x1x6 x2x4 x3x7 x5x8 + x1x4 x2x6 x3x7 x5x8 +<br />

x1x7 x2x3 x4x6 x5x8 + x1x3 x2x7 x4x6 x5x8 + x1x2 x3x7 x4x6 x5x8 +<br />

x1x6 x2x3 x4x7 x5x8 + x1x3 x2x6 x4x7 x5x8 + x1x2 x3x6 x4x7 x5x8 +<br />

x1x8 x2x5 x3x4 x6x7 + x1x5 x2x8 x3x4 x6x7 + x1x8 x2x4 x3x5 x6x7 +<br />

x1x4 x2x8 x3x5 x6x7 + x1x5 x2x4 x3x8 x6x7 + x1x4 x2x5 x3x8 x6x7 +<br />

x1x8 x2x3 x4x5 x6x7 + x1x3 x2x8 x4x5 x6x7 + x1x2 x3x8 x4x5 x6x7 +<br />

x1x5 x2x3 x4x8 x6x7 + x1x3 x2x5 x4x8 x6x7 + x1x2 x3x5 x4x8 x6x7 +<br />

x1x7 x2x5 x3x4 x6x8 + x1x5 x2x7 x3x4 x6x8 + x1x7 x2x4 x3x5 x6x8 +<br />

x1x4 x2x7 x3x5 x6x8 + x1x5 x2x4 x3x7 x6x8 + x1x4 x2x5 x3x7 x6x8 +<br />

x1x7 x2x3 x4x5 x6x8 + x1x3 x2x7 x4x5 x6x8 + x1x2 x3x7 x4x5 x6x8 +<br />

x1x5 x2x3 x4x7 x6x8 + x1x3 x2x5 x4x7 x6x8 + x1x2 x3x5 x4x7 x6x8 +<br />

x1x6 x2x5 x3x4 x7x8 + x1x5 x2x6 x3x4 x7x8 + x1x6 x2x4 x3x5 x7x8 +<br />

x1x4 x2x6 x3x5 x7x8 + x1x5 x2x4 x3x6 x7x8 + x1x4 x2x5 x3x6 x7x8 +<br />

x1x6 x2x3 x4x5 x7x8 + x1x3 x2x6 x4x5 x7x8 + x1x2 x3x6 x4x5 x7x8 +<br />

x1x5 x2x3 x4x6 x7x8 + x1x3 x2x5 x4x6 x7x8 + x1x2 x3x5 x4x6 x7x8<br />

Which are 96 terms. Now if we notice that every term which has a x(2k + 1)x2k will<br />

only contribute to κ 0 then we can get rid of many more terms. And these will be the<br />

terms we are left <strong>with</strong>:<br />

x1x8 x2x7 x3x6 x4x5 + x1x7 x2x8 x3x6 x4x5 + x1x8 x2x6 x3x7 x4x5 +<br />

x1x6 x2x8 x3x7 x4x5 + x1x7 x2x6 x3x8 x4x5 + x1x6 x2x7 x3x8 x4x5 +<br />

x1x8 x2x7 x3x5 x4x6 + x1x7 x2x8 x3x5 x4x6 + x1x8 x2x5 x3x7 x4x6 +<br />

x1x5 x2x8 x3x7 x4x6 + x1x7 x2x5 x3x8 x4x6 + x1x5 x2x7 x3x8 x4x6 +<br />

x1x8 x2x6 x3x5 x4x7 + x1x6 x2x8 x3x5 x4x7 + x1x8 x2x5 x3x6 x4x7 +<br />

x1x5 x2x8 x3x6 x4x7 + x1x6 x2x5 x3x8 x4x7 + x1x5 x2x6 x3x8 x4x7 +<br />

x1x7 x2x6 x3x5 x4x8 + x1x6 x2x7 x3x5 x4x8 + x1x7 x2x5 x3x6 x4x8 +<br />

x1x5 x2x7 x3x6 x4x8 + x1x6 x2x5 x3x7 x4x8 + x1x5 x2x6 x3x7 x4x8 +<br />

x1x8 x2x4 x3x6 x5x7 + x1x4 x2x8 x3x6 x5x7 + x1x6 x2x4 x3x8 x5x7 +<br />

x1x4 x2x6 x3x8 x5x7 + x1x8 x2x3 x4x6 x5x7 + x1x3 x2x8 x4x6 x5x7 +<br />

x1x6 x2x3 x4x8 x5x7 + x1x3 x2x6 x4x8 x5x7 + x1x7 x2x4 x3x6 x5x8 +<br />

x1x4 x2x7 x3x6 x5x8 + x1x6 x2x4 x3x7 x5x8 + x1x4 x2x6 x3x7 x5x8 +<br />

x1x7 x2x3 x4x6 x5x8 + x1x3 x2x7 x4x6 x5x8 + x1x6 x2x3 x4x7 x5x8 +<br />

x1x3 x2x6 x4x7 x5x8 + x1x8 x2x4 x3x5 x6x7 + x1x4 x2x8 x3x5 x6x7 +<br />

x1x5 x2x4 x3x8 x6x7 + x1x4 x2x5 x3x8 x6x7 + x1x8 x2x3 x4x5 x6x7 +<br />

x1x3 x2x8 x4x5 x6x7 + x1x5 x2x3 x4x8 x6x7 + x1x3 x2x5 x4x8 x6x7 +


B.0 142<br />

x1x7 x2x4 x3x5 x6x8 + x1x4 x2x7 x3x5 x6x8 + x1x5 x2x4 x3x7 x6x8 +<br />

x1x4 x2x5 x3x7 x6x8 + x1x7 x2x3 x4x5 x6x8 + x1x3 x2x7 x4x5 x6x8 +<br />

x1x5 x2x3 x4x7 x6x8 + x1x3 x2x5 x4x7 x6x8<br />

Which are 56 terms.<br />

Now if we remember that<br />

x1 = R1q1 (B.2)<br />

x2 = R1q2<br />

x3 = R2q1<br />

x4 = R2q2<br />

x5 = R3q3<br />

x6 = R3q4<br />

x7 = R4q3<br />

x8 = R4q4<br />

The above terms can be written in the following way.<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+


B.0 143<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

<br />

It is easier to use the simpler expression for BiPS which we derived in A. Using this<br />

simplified form of BiPS, we can rewrite the kernel of CV in the following form:<br />

+<br />

+<br />

+<br />

+<br />

+


B.0 144<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+


B.0 145<br />

+<br />

+<br />

+<br />

+<br />

+<br />

+<br />

<br />

All of these temrs are somehow similar. And all of them have the same integrals. So if we<br />

do some of them we can find a shortcut to the answer of integration of each term. This<br />

was explained in A and can be seen that having known that, the rest of our calculation<br />

is more like a book keeping task. We should very carefully keep track of indices. I wrote<br />

a FORTRAN90 program (!) to do that. It reads the above expression and outputs the<br />

expression below in LATEX format,<br />

C l1 C l3 C l1 C l3 C lM<br />

l 1 −m 1 l 3 −m 3<br />

C lM<br />

l 1 m 5 l 3 m 7<br />

C lM ′<br />

l 1 m 5 l 3 m 7<br />

C lM ′<br />

l 1 −m 1 l 3 −m 3<br />

+<br />

C l1 C l1 C l1 C l1 C lM<br />

l 1 −m 1 l 1 −m 3<br />

C lM<br />

l 1 m 5 l 1 m 7<br />

C lM ′<br />

l 1 m 7 l 1 m 5<br />

C lM ′<br />

l 1 −m 1 l 1 −m 3<br />

+<br />

C l1 C l1 C l1 C l1 C lM<br />

l 1 −m 1 l 1 −m 3<br />

C lM<br />

l 1 m 5 l 1 m 7<br />

C lM ′<br />

l 1 m 3 l 1 m 7<br />

C lM ′<br />

l 1 −m 1 l 1 −m 5<br />

+<br />

C l1 C l1 C l1 C l1 C lM<br />

l 1 −m 1 l 1 −m 3<br />

C lM<br />

l 1 m 5 l 1 m 7<br />

C lM ′<br />

l 1 m 1 l 1 m 7<br />

C lM ′<br />

l 1 −m 3 l 1 −m 5<br />

+<br />

C l1 C l1 C l1 C l1 C lM<br />

l 1 −m 1 l 1 −m 3<br />

C lM<br />

l 1 m 5 l 1 m 7<br />

C lM ′<br />

l 1 m 7 l 1 m 3<br />

C lM ′<br />

l 1 −m 1 l 1 −m 5<br />

+<br />

C l1 C l3 C l1 C l3 C lM<br />

l 1 −m 1 l 3 −m 3<br />

C lM<br />

l 1 m 5 l 3 m 7<br />

C lM ′<br />

l 3 m 7 l 1 m 1<br />

C lM ′<br />

l 3 −m 3 l 1 −m 5<br />

+<br />

C l1 C l1 C l1 C l1 C lM<br />

l 1 −m 1 l 1 −m 3<br />

C lM<br />

l 1 m 5 l 1 m 7<br />

C lM ′<br />

l 1 m 7 l 1 m 5<br />

C lM ′<br />

l 1 −m 1 l 1 −m 3<br />

+<br />

C l1 C l3 C l1 C l3 C lM<br />

l 1 −m 1 l 3 −m 3<br />

C lM<br />

l 1 m 5 l 3 m 7<br />

C lM ′<br />

l 1 m 5 l 3 m 7<br />

C lM ′<br />

l 1 −m 1 l 3 −m 3<br />

+<br />

C l1 C l1 C l1 C l1 C lM<br />

l 1 −m 1 l 1 −m 3<br />

C lM<br />

l 1 m 5 l 1 m 7<br />

C lM ′<br />

l 1 m 7 l 1 m 3<br />

C lM ′<br />

l 1 −m 1 l 1 −m 5<br />

+<br />

C l1 C l3 C l1 C l3 C lM<br />

l 1 −m 1 l 3 −m 3<br />

C lM<br />

l 1 m 5 l 3 m 7<br />

C lM ′<br />

l 3 m 7 l 1 m 1<br />

C lM ′<br />

l 3 −m 3 l 1 −m 5<br />

+<br />

C l1 C l1 C l1 C l1 C lM<br />

l 1 −m 1 l 1 −m 3<br />

C lM<br />

l 1 m 5 l 1 m 7<br />

C lM ′<br />

l 1 m 3 l 1 m 7<br />

C lM ′<br />

l 1 −m 1 l 1 −m 5<br />

+<br />

C l1 C l1 C l1 C l1 C lM<br />

l 1 −m 1 l 1 −m 3<br />

C lM<br />

l 1 m 5 l 1 m 7<br />

C lM ′<br />

l 1 m 1 l 1 m 7<br />

C lM ′<br />

l 1 −m 3 l 1 −m 5<br />

+<br />

C l1 C l1 C l1 C l1 C lM<br />

l 1 −m 1 l 1 −m 3<br />

C lM<br />

l 1 m 5 l 1 m 7<br />

C lM ′<br />

l 1 m 3 l 1 m 5<br />

C lM ′<br />

l 1 −m 1 l 1 −m 7<br />

+<br />

C l1 C l1 C l1 C l1 C lM<br />

l 1 −m 1 l 1 −m 3<br />

C lM<br />

l 1 m 5 l 1 m 7<br />

C lM ′<br />

l 1 m 1 l 1 m 5<br />

C lM ′<br />

l 1 −m 3 l 1 −m 7<br />

+<br />

C l1 C l3 C l1 C l3 C lM<br />

l 1 −m 1 l 3 −m 3<br />

C lM<br />

l 1 m 5 l 3 m 7<br />

C lM ′<br />

l 1 m 5 l 3 m 3<br />

C lM ′<br />

l 1 −m 1 l 3 −m 7<br />

+<br />

C l1 C l1 C l1 C l1 C lM<br />

l 1 −m 1 l 1 −m 3<br />

C lM<br />

l 1 m 5 l 1 m 7<br />

C lM ′<br />

l 1 m 5 l 1 m 1<br />

C lM ′<br />

l 1 −m 3 l 1 −m 7<br />

+


B.0 146<br />

C l1 C l3 C l1 C l3 C lM<br />

l 1 −m 1 l 3 −m 3<br />

C lM<br />

l 1 m 5 l 3 m 7<br />

C lM ′<br />

l 1 m 1 l 3 m 3<br />

C lM ′<br />

l 1 −m 5 l 3 −m 7<br />

+<br />

C l1 C l1 C l1 C l1 C lM<br />

l 1 −m 1 l 1 −m 3<br />

C lM<br />

l 1 m 5 l 1 m 7<br />

C lM ′<br />

l 1 m 3 l 1 m 1<br />

C lM ′<br />

l 1 −m 5 l 1 −m 7<br />

+<br />

C l1 C l3 C l1 C l3 C lM<br />

l 1 −m 1 l 3 −m 3<br />

C lM<br />

l 1 m 5 l 3 m 7<br />

C lM ′<br />

l 1 m 5 l 3 m 3<br />

C lM ′<br />

l 1 −m 1 l 3 −m 7<br />

+<br />

C l1 C l1 C l1 C l1 C lM<br />

l 1 −m 1 l 1 −m 3<br />

C lM<br />

l 1 m 5 l 1 m 7<br />

C lM ′<br />

l 1 m 5 l 1 m 1<br />

C lM ′<br />

l 1 −m 3 l 1 −m 7<br />

+<br />

C l1 C l1 C l1 C l1 C lM<br />

l 1 −m 1 l 1 −m 3<br />

C lM<br />

l 1 m 5 l 1 m 7<br />

C lM ′<br />

l 1 m 3 l 1 m 5<br />

C lM ′<br />

l 1 −m 1 l 1 −m 7<br />

+<br />

C l1 C l1 C l1 C l1 C lM<br />

l 1 −m 1 l 1 −m 3<br />

C lM<br />

l 1 m 5 l 1 m 7<br />

C lM ′<br />

l 1 m 1 l 1 m 5<br />

C lM ′<br />

l 1 −m 3 l 1 −m 7<br />

+<br />

C l1 C l1 C l1 C l1 C lM<br />

l 1 −m 1 l 1 −m 3<br />

C lM<br />

l 1 m 5 l 1 m 7<br />

C lM ′<br />

l 1 m 3 l 1 m 1<br />

C lM ′<br />

l 1 −m 5 l 1 −m 7<br />

+<br />

C l1 C l3 C l1 C l3 C lM<br />

l 1 −m 1 l 3 −m 3<br />

C lM<br />

l 1 m 5 l 3 m 7<br />

C lM ′<br />

l 1 m 1 l 3 m 3<br />

C lM ′<br />

l 1 −m 5 l 3 −m 7<br />

+<br />

C l1 C l3 C l1 C l7 C lM<br />

l 3 −m 3 l 1 −m 1<br />

C lM<br />

l 3 −m 3 l 1 m 5<br />

C lM ′<br />

l 1 m 5 l 7 −m 7<br />

C lM ′<br />

l 1 −m 1 l 7 −m 7<br />

+<br />

C l1 C l1 C l1 C l7 C lM<br />

l 1 −m 3 l 1 −m 1<br />

C lM<br />

l 1 −m 1 l 1 m 5<br />

C lM ′<br />

l 1 m 5 l 7 −m 7<br />

C lM ′<br />

l 1 −m 3 l 7 −m 7<br />

+<br />

C l1 C l3 C l1 C l7 C lM<br />

l 3 −m 3 l 1 −m 1<br />

C lM<br />

l 3 −m 3 l 1 m 5<br />

C lM ′<br />

l 1 m 1 l 7 −m 7<br />

C lM ′<br />

l 1 −m 5 l 7 −m 7<br />

+<br />

C l1 C l1 C l1 C l7 C lM<br />

l 1 m 10 l 1 −m 1<br />

C lM<br />

l 1 −m 1 l 1 m 5<br />

C lM ′<br />

l 1 −m 2 l 7 −m 7<br />

C lM ′<br />

l 1 −m 5 l 7 −m 7<br />

+<br />

C l1 C l1 C l1 C l7 C lM<br />

l 1 −m 1 l 1 −m 3<br />

C lM<br />

l 1 −m 3 l 1 m 5<br />

C lM ′<br />

l 1 m 5 l 7 −m 7<br />

C lM ′<br />

l 1 −m 1 l 7 −m 7<br />

+<br />

C l1 C l3 C l3 C l7 C lM<br />

l 1 −m 1 l 3 −m 3<br />

C lM<br />

l 1 −m 1 l 3 m 5<br />

C lM ′<br />

l 3 m 5 l 7 −m 7<br />

C lM ′<br />

l 3 −m 3 l 7 −m 7<br />

+<br />

C l1 C l1 C l1 C l7 C lM<br />

l 1 −m 1 l 1 −m 3<br />

C lM<br />

l 1 −m 3 l 1 m 5<br />

C lM ′<br />

l 1 m 1 l 7 −m 7<br />

C lM ′<br />

l 1 −m 5 l 7 −m 7<br />

+<br />

C l1 C l3 C l3 C l7 C lM<br />

l 1 −m 1 l 3 −m 3<br />

C lM<br />

l 1 −m 1 l 3 m 5<br />

C lM ′<br />

l 3 m 3 l 7 −m 7<br />

C lM ′<br />

l 3 −m 5 l 7 −m 7<br />

+<br />

C l1 C l3 C l1 C l1 C lM<br />

l 3 −m 3 l 1 −m 1<br />

C lM<br />

l 3 −m 3 l 1 m 5<br />

C lM ′<br />

l 1 −m 7 l 1 m 5<br />

C lM ′<br />

l 1 −m 1 l 1 −m 7<br />

+<br />

C l1 C l1 C l1 C l1 C lM<br />

l 1 m 10 l 1 −m 1<br />

C lM<br />

l 1 −m 1 l 1 m 5<br />

C lM ′<br />

l 1 −m 7 l 1 m 5<br />

C lM ′<br />

l 1 m 2 l 1 −m 7<br />

+<br />

C l1 C l3 C l1 C l1 C lM<br />

l 3 −m 3 l 1 −m 1<br />

C lM<br />

l 3 −m 3 l 1 m 5<br />

C lM ′<br />

l 1 −m 7 l 1 m 1<br />

C lM ′<br />

l 1 −m 5 l 1 −m 7<br />

+<br />

C l1 C l1 C l1 C l1 C lM<br />

l 1 m 10 l 1 −m 1<br />

C lM<br />

l 1 −m 1 l 1 m 5<br />

C lM ′<br />

l 1 −m 7 l 1 −m 2<br />

C lM ′<br />

l 1 −m 5 l 1 −m 7<br />

+<br />

C l1 C l1 C l1 C l1 C lM<br />

l 1 −m 1 l 1 −m 3<br />

C lM<br />

l 1 −m 3 l 1 m 5<br />

C lM ′<br />

l 1 −m 7 l 1 m 5<br />

C lM ′<br />

l 1 −m 1 l 1 −m 7<br />

+<br />

C l1 C l3 C l3 C l3 C lM<br />

l 1 −m 1 l 3 −m 3<br />

C lM<br />

l 1 −m 1 l 3 m 5<br />

C lM ′<br />

l 3 −m 7 l 3 m 5<br />

C lM ′<br />

l 3 −m 3 l 3 −m 7<br />

+<br />

C l1 C l1 C l1 C l1 C lM<br />

l 1 −m 1 l 1 −m 3<br />

C lM<br />

l 1 −m 3 l 1 m 5<br />

C lM ′<br />

l 1 −m 7 l 1 m 1<br />

C lM ′<br />

l 1 −m 5 l 1 −m 7<br />

+<br />

C l1 C l3 C l3 C l3 C lM<br />

l 1 −m 1 l 3 −m 3<br />

C lM<br />

l 1 −m 1 l 3 m 5<br />

C lM ′<br />

l 3 −m 7 l 3 m 3<br />

C lM ′<br />

l 3 −m 5 l 3 −m 7<br />

+<br />

C l1 C l3 C l1 C l1 C lM<br />

l 3 −m 3 l 1 −m 1<br />

C lM<br />

l 3 −m 3 l 1 m 5<br />

C lM ′<br />

l 1 −m 7 l 1 m 5<br />

C lM ′<br />

l 1 −m 1 l 1 −m 7<br />

+<br />

C l1 C l1 C l1 C l1 C lM<br />

l 1 −m 3 l 1 −m 1<br />

C lM<br />

l 1 −m 1 l 1 m 5<br />

C lM ′<br />

l 1 −m 7 l 1 m 5<br />

C lM ′<br />

l 1 −m 3 l 1 −m 7<br />

+<br />

C l1 C l3 C l1 C l1 C lM<br />

l 3 −m 3 l 1 −m 1<br />

C lM<br />

l 3 −m 3 l 1 m 5<br />

C lM ′<br />

l 1 −m 7 l 1 m 1<br />

C lM ′<br />

l 1 −m 5 l 1 −m 7<br />

+


B.0 147<br />

C l1 C l1 C l1 C l1 C lM<br />

l 1 −m 3 l 1 −m 1<br />

C lM<br />

l 1 −m 1 l 1 m 5<br />

C lM ′<br />

l 1 −m 7 l 1 m 3<br />

C lM ′<br />

l 1 −m 5 l 1 −m 7<br />

+<br />

C l1 C l1 C l1 C l1 C lM<br />

l 1 −m 1 l 1 −m 3<br />

C lM<br />

l 1 −m 3 l 1 m 5<br />

C lM ′<br />

l 1 −m 7 l 1 m 5<br />

C lM ′<br />

l 1 −m 1 l 1 −m 7<br />

+<br />

C l1 C l3 C l3 C l3 C lM<br />

l 1 −m 1 l 3 −m 3<br />

C lM<br />

l 1 −m 1 l 3 m 5<br />

C lM ′<br />

l 3 −m 7 l 3 m 5<br />

C lM ′<br />

l 3 −m 3 l 3 −m 7<br />

+<br />

C l1 C l1 C l1 C l1 C lM<br />

l 1 −m 1 l 1 −m 3<br />

C lM<br />

l 1 −m 3 l 1 m 5<br />

C lM ′<br />

l 1 −m 7 l 1 m 1<br />

C lM ′<br />

l 1 −m 5 l 1 −m 7<br />

+<br />

C l1 C l3 C l3 C l3 C lM<br />

l 1 −m 1 l 3 −m 3<br />

C lM<br />

l 1 −m 1 l 3 m 5<br />

C lM ′<br />

l 3 −m 7 l 3 m 3<br />

C lM ′<br />

l 3 −m 5 l 3 −m 7<br />

+<br />

C l1 C l3 C l1 C l7 C lM<br />

l 3 −m 3 l 1 −m 1<br />

C lM<br />

l 3 −m 3 l 1 m 5<br />

C lM ′<br />

l 1 m 5 l 7 −m 7<br />

C lM ′<br />

l 1 −m 1 l 7 −m 7<br />

+<br />

C l1 C l1 C l1 C l7 C lM<br />

l 1 m 10 l 1 −m 1<br />

C lM<br />

l 1 −m 1 l 1 m 5<br />

C lM ′<br />

l 1 m 5 l 7 −m 7<br />

C lM ′<br />

l 1 m 2 l 7 −m 7<br />

+<br />

C l1 C l3 C l1 C l7 C lM<br />

l 3 −m 3 l 1 −m 1<br />

C lM<br />

l 3 −m 3 l 1 m 5<br />

C lM ′<br />

l 1 m 1 l 7 −m 7<br />

C lM ′<br />

l 1 −m 5 l 7 −m 7<br />

+<br />

C l1 C l1 C l1 C l7 C lM<br />

l 1 m 10 l 1 −m 1<br />

C lM<br />

l 1 −m 1 l 1 m 5<br />

C lM ′<br />

l 1 −m 2 l 7 −m 7<br />

C lM ′<br />

l 1 −m 5 l 7 −m 7<br />

+<br />

C l1 C l1 C l1 C l7 C lM<br />

l 1 −m 1 l 1 −m 3<br />

C lM<br />

l 1 −m 3 l 1 m 5<br />

C lM ′<br />

l 1 m 5 l 7 −m 7<br />

C lM ′<br />

l 1 −m 1 l 7 −m 7<br />

+<br />

C l1 C l3 C l3 C l7 C lM<br />

l 1 −m 1 l 3 −m 3<br />

C lM<br />

l 1 −m 1 l 3 m 5<br />

C lM ′<br />

l 3 m 5 l 7 −m 7<br />

C lM ′<br />

l 3 −m 3 l 7 −m 7<br />

+<br />

C l1 C l1 C l1 C l7 C lM<br />

l 1 −m 1 l 1 −m 3<br />

C lM<br />

l 1 −m 3 l 1 m 5<br />

C lM ′<br />

l 1 m 1 l 7 −m 7<br />

C lM ′<br />

l 1 −m 5 l 7 −m 7<br />

+<br />

C l1 C l3 C l3 C l7 C lM<br />

l 1 −m 1 l 3 −m 3<br />

C lM<br />

l 1 −m 1 l 3 m 5<br />

C lM ′<br />

l 3 m 3 l 7 −m 7<br />

C lM ′<br />

l 3 −m 5 l 7 −m 7<br />

(B.3)<br />

This can be then simplified much further to get the final expression for cosmic<br />

variance of BiPS as it is explained in Appendix A.


AppendixC<br />

Search for Planar Symmetries in <strong>CMB</strong><br />

Ansiotropy Maps<br />

Flat spaces have planar symmetries which can be used to detect them. These symmetries<br />

as pointed out by Starobinsky<br />

(1993), emerge if the size of the fundamental<br />

domain is smaller than the diameter of the surface of last scattering.<br />

For example,<br />

imagine a case where the fundamental domain is a cuboid and only one of its cell sizes<br />

is smaller than R SLS . In this case the large scale temperature fluctuations have a symmetry<br />

plane which is perpendicular to the compact direction. One can then look for<br />

these planar symmetries as signatures of compact flat spaces. We perform this search<br />

for symmetries using the statistics originally proposed in de Oliveira-Costa & Smoot<br />

(1995) and later made easier by de Oliveira-Costa et al. (2003). This is known as<br />

S-statistics and is done by computing the function S(ˆn i ) defined by<br />

S(ˆn i ) = 1<br />

N<br />

∑ pix ( ∆T<br />

N pix T (ˆn j) − ∆T ) 2<br />

T (ˆn ij)<br />

(C.1)<br />

j=1<br />

where N pix is the number of pixels in the <strong>CMB</strong> map and ˆn ij denotes the reflection of ˆn j<br />

in the plane whose normal is ˆn i , i.e.,<br />

ˆn ij = ˆn j − 2(ˆn i · ˆn j )ˆn i<br />

(C.2)<br />

The original S-statistics proposed by de Oliveira-Costa & Smoot (1995) was slightly<br />

different and had a division by the r.m.s. errors associated <strong>with</strong> the pixels in the form<br />

148


C.0 149<br />

of<br />

S(ˆn i ) = 1<br />

N<br />

∑ pix<br />

N pix<br />

j=1<br />

( ∆T<br />

(ˆn T j) − ∆T<br />

σ 2 (ˆn j ) + σ 2 (ˆn ij )<br />

(ˆn T ij) ) 2<br />

. (C.3)<br />

But we used the one given in eqn. (C.1) because at that time we did not know the r.m.s.<br />

pixel noise of the WMAP data.<br />

S(ˆn i ) itself can be viewed as a map, and is a measure of how much reflection symmetry<br />

there is in the mirror plane perpendicular to S(ˆn i ). This resultant map is referred<br />

to as an S-map. This is a very useful visualization tool to search for symmetries and<br />

works in the following way.<br />

(a)<br />

(b)<br />

As it is seen in eqn. C.1, the S(ˆn i ) is a sum over non-negative quantities.<br />

These quantities are the difference between ∆T/T at every pixel and it’s<br />

reflection <strong>with</strong> respect to an axis. At the axis of symmetry, ∆T/T at each<br />

pixel is similar to the ∆T/T at the position of reflection of this point <strong>with</strong><br />

respect to that axis. And hence the minimum of the S-map, denoted as S 0 ,<br />

which is visualized as the coldest spot of the map is pointing to the axis of<br />

symmetry.<br />

The distribution of S 0 values changes <strong>with</strong> the size of the fundamental domain.<br />

The smaller the fundamental domain, the smaller the value of S 0<br />

would be.<br />

(c)<br />

The value of S 0 depends on the size of the fundamental domain and is independent<br />

of its orientation.<br />

The S-statistics will then be best done by obtaining S-map and looking for the coldest<br />

spots on it. We computed the S-map for ILC map. This map as it is shown in Fig. C.1<br />

has some cold spots, major ones are on the poles and two long ones (spots appear in<br />

pairs because the S-map has been built symmetrically). In Fig. C.1 we show the S-map<br />

obtained from a random simulation of the <strong>CMB</strong> sky <strong>with</strong> the best fit C l of WMAP. The<br />

patterns on these two S-maps are clearly different. But what is important here, is how<br />

cold the coldest spots, S 0 , in the ILC S-map are. To answer this question, we degrade<br />

the ILC map to the HEALPix resolution of N side = 32 and compute the S 0 in that. We<br />

will then make 500 random realizations of the <strong>CMB</strong> sky from the best fit C l of WMAP


C.0 150<br />

and compute the S 0 in each of them. The question of how cold is the coldest spot in<br />

ILC S-map would then be equivalent to in the random realizations of a normal <strong>CMB</strong><br />

sky, how many times do I get S 0 values bigger than that of S 0 of ILC map. If ILC<br />

represents the <strong>CMB</strong> in a small compact flat space, its S 0 should be smaller than most<br />

of the S 0 values obtained from the random realizations. The result of this test is shown<br />

in the left panel of Fig. C.2. Which means that S 0 of ILC is not low at all. And this<br />

rules out the possibility of small spaces. The above test was done by de Oliveira-Costa<br />

et al. (1996) for the 4 year COBE/DMR data and by de Oliveira-Costa et al. (2003)<br />

for the WMAP data. However, the quadrupole in the first year WMAP data has been<br />

the source of many debates. To make sure that we have not been misled by that, we<br />

do the above test once again, this time by excluding the quadrupole from ILC map as<br />

well as the simulated maps. The result is shown in right panel of Fig. C.2 and is still<br />

consistent <strong>with</strong> a not-so-low S 0 of the ILC map.


C.0 151<br />

Figure C.1 SMAP of ILC map (top) and of a random realization of the <strong>CMB</strong> sky obtained<br />

from the best fit C l of the WMAP data (bottom). Units are arbitrary and figures are<br />

presented for a shape comparison of the two maps.


C.0 152


AppendixD<br />

Symmetries of the Space<br />

By symmetries of the space, we mean those transformations of the space into itself<br />

under which metric and every other physical and geometrical properties of the space<br />

remain invariant. Here we only consider continuous symmetries. These symmetries as we<br />

shal see, form continuous groups of symmetries and are generated by vector fields which<br />

are generators of these groups. In the following section we will study the mathematical<br />

preliminaries of the notions used in Chapter 5.<br />

Mathematical preliminaries<br />

Let M be an m-dimensional differentiable manifold and X be a vector field in M.<br />

An integral curve x(t) of X is a curve in M, whose tangent vector at x(t) is X| x<br />

dx µ (t)<br />

dt<br />

= X µ (x(t)) (D.1)<br />

Let σ(t, x 0 ) be an integral curve of X which passes a point x 0 at t = 0 and denote the<br />

coordinate by σ µ (t, x 0 ). Eqn. (D.1) then becomes<br />

<strong>with</strong> the initial condition<br />

d<br />

dt σµ (t, x 0 ) = X µ (σ(t, x 0 ))<br />

σ µ (0, x 0 ) = x µ 0 .<br />

(D.2)<br />

(D.3)<br />

The map σ : R × M ↦−→ M is called a flow generated by X ∈ X(M) 1 . For fixed t ∈ R<br />

a flow is a diffeomorphism from M to M, denoted by σ t : M ↦−→ M and is made into a<br />

1 X(M) denotes the set of vector fields on M.<br />

153


D.0 154<br />

commutative group by the folloing rules<br />

1. σ t (σ s (x)) = σ t+s (x),<br />

2. σ 0 = the identity map,<br />

3. σ −t = (σ t ) −1 . This group is called the one-parameter group of transformations.<br />

Diffeomorphism: A C ∞ map is called a diffeomorphism if it is one-to-one and onto<br />

(and hence invertible) and its inverse is C ∞ too. Let (M, g) be a (pseudo-) Riemannian<br />

manifold. A diffeomorphism f : M → M is an isometry if it preserves the metric<br />

f ∗ g f(p) = g p .<br />

(D.4)<br />

This means that f is an isometry if for each X, Y ∈ T p M we have g f(p) (f ∗ X, f ∗ Y ) =<br />

g p (X, Y ). In components, this condition becomes<br />

∂y α<br />

∂x µ ∂y β<br />

∂x ν g αβ(f(p)) = g µν (p),<br />

(D.5)<br />

where x and y are coordinates of p and f(p) respectively. The isometries of a space of<br />

dimension n form a group, because the identity map, the composition of two isometries<br />

and the inverse of an isometry are all isometries too. For example in R n , the Euclidean<br />

group E n is the isometry group, that is f : x → Ax + T , where A ∈ SO(n) are rotations<br />

and T ∈ R n causes the translations.<br />

Let (M, g) be a Riemannian manifold and X ∈ X(M). If an infinitesimal displacement<br />

given by ɛX generates an isometry (ɛ is an infinitesimal parameter), the vector<br />

field X is called a Killing vector fields. The coordinates x µ of a point p ∈ M change<br />

to x µ + ɛX µ (p) under this displacement 2 . If f : x µ ↦−→ x µ + ɛX µ is an isometry, it<br />

satisfies (D.5),<br />

∂(x α + ɛX α )<br />

∂x µ ∂(x β + ɛX β )<br />

∂x ν g αβ (x + ɛX) = g µν (x) (D.6)<br />

The above expression tells us that g µν and X µ satisfy the Killing equation<br />

X η ∂ η g µν + ∂ µ X η g ην + ∂ ν X η g µη = 0.<br />

(D.7)<br />

2 σ µ ɛ (x) = σ µ (ɛ, x) = x µ + ɛX µ (x) is a one-parameter group of transformations and the vector field<br />

X is called the infinitesimal generator of this transformation.


D.0 155<br />

And from the definition of the Lie derivative, this is written as<br />

(L X g) µν = 0. (D.8)<br />

This has a nice interpretation. Let φ t : M ↦−→ M be a one-parameter group of transformations<br />

which generate the Killing vector field X. Eqn. (D.8 then shows that the local<br />

geometry does not change as we move along φ t . In this sense, the Killing vector fields<br />

represent the direction of the symmetry of a manifold.<br />

A set of Killing vector fields is said to be linearly dependent if one of them is expressed<br />

as a linear combination of others <strong>with</strong> constant coefficients 3 . The maximum number of<br />

symmetries is related to dimM and in an m-dimensional manifold is m(m+1)/2. Spaces<br />

which admit m(m+1)/2 Killing vector fields are called maximally symmetric spaces.<br />

Let X a and X b be two Killing vector fields. It can be seen that<br />

(a)<br />

a linear combination αX a + βX b is a killing vector field (α, β ∈ R),<br />

(b) the commutator [X a , X b ] is a Killing vector field 4 .<br />

Thus all Killing vector fields form a Lie algebra of the symmetric operations on the<br />

manifold M, <strong>with</strong> structure constants Cab<br />

c<br />

[X a , X b ] = Cab c X c (D.9)<br />

where a, b, c = 1, 2, ..., r and r ≤ m(m + 1)/2. Structure constants satisfy<br />

(a)<br />

(b)<br />

Skew-symmetry<br />

Jacobi identity<br />

C c ab = −C c ba<br />

C a e[b Ce cd] = 0<br />

(D.10)<br />

(D.11)<br />

These are the integrability conditions that must be satisfied in order that the Lie algebra<br />

exist in a consistent way. The transformations generated by the Lie algebra form a Lie<br />

group of the same dimension.<br />

3 There may be more Killing vector fields than the dimension of the manifold. The number of<br />

independent symmetries has no direct connection <strong>with</strong> the dimension of M.<br />

4 Because [X a , X b ] g = L Xa (L Xb g) − L Xb (L Xa g) = 0


D.0 156<br />

In the case of our interest, a Lie group often appears as a set of transformations<br />

acting on a manifold. We need to study the action of a Lie group G on a manifold M<br />

in more details. Let G be a Lie group and M a manifold. The action of G on M is a<br />

differentiable map σ : G × M ↦−→ M which satisfies the conditions<br />

(i) σ(e, p) = p for any p ∈ M<br />

(ii)<br />

σ(g 1 , σ(g 2 , p)) = σ(g 1 g 2 , p).<br />

The action σ is said to be transitive if, for any p 1 , p 2 ∈ M, there exists an element<br />

g ∈ G such that σ(g, p 1 ) = p 2 . Which means it can move any point of M into any other<br />

point of M. Given a point p ∈ M, the action of G on p takes p to various points in M.<br />

The orbit of p is the set of all points into which p can be moved by the action of the<br />

isometries of a space. This is<br />

Gp = {σ(g, p)|g ∈ G}.<br />

(D.12)<br />

Any orbit Gp is clearly a subset of M and obviously the action of G on any orbit Gp is<br />

transitive. Orbits are in fact the largest surfaces through each point on which the group<br />

is transitive. If the action of G on M is transitive, the orbit of any p ∈ M is M itself.<br />

So, the maximum dimension of orbits is dimM.<br />

dim orbits = s, where in a space of dimension n (dimM = n), s ≤ n.<br />

A subgroup H(p) of a Lie group G that acts on M is called an isotropy group if<br />

it leaves the point p ∈ M fixed,<br />

H(p) = {g ∈ G|σ(g, p) = p}<br />

(D.13)<br />

H(p) is also called the little group or stabilizer. We have<br />

dim of isotropy group = q, where q ≤ 1 n(n − 1).<br />

2<br />

Interestingly, the cosets of the isotropy group correspond to the elements in the orbit,<br />

G p ∼ G/H(p)<br />

(D.14)<br />

G/H has a very interesting property indeed. There is a theorem which says Nakahara<br />

(1984): for any subgroup H of a Lie group G, the coset space G/H admits a differentiable


D.0 157<br />

structure and becomes a manifold called homogeneous space. Dimension of this coset<br />

space is given by<br />

dimG/H = dimG − dimH<br />

(D.15)<br />

In our case, G is a Lie group which acts on a manifold M transitively, and H(p) is an<br />

isotropy group of p ∈ M. This theorem tells us that H(p) is a Lie group and the coset<br />

G/H(p) is a homogeneous space.<br />

The dimension r of the group of symmetries of the space (group of isometries), is<br />

then given by<br />

r = s + q =⇒ 0 ≤ r ≤ n + n 2 (n − 1) = 1 n(n + 1)<br />

2 (D.16)<br />

which is what we had seen before. The above relation can be viewed as<br />

dim group of isometries = dim group of rotational symmetries + dim group of translational<br />

symmetries.<br />

If q = 0 then r = s which means that dimension of group of isometries is just enough<br />

to move each point in an orbit into any other point. This is called a simply transitive<br />

group. There is no continuous isotropy group in this case.


AppendixE<br />

Bianchi Models: Bianchi’s Original<br />

Classification<br />

In his historical paper On the 3-dimensional spaces which admit a continous group<br />

of motions 1 , Luigi Bianchi begins <strong>with</strong> finite-dimensional continuous Lie group G r generated<br />

by r infinitesimal transformations X 1 , X 2 , ..., X r of an n-dimensional Riemannian<br />

manifold. Then his problem of determining which spaces possess a continuous group of<br />

motions gets reduced to the classification of all possible forms of metrics which possess<br />

a Lie group G r = {X 1 , · · · , X n } which transforms the metric into itself. Lie’s classification<br />

of equivalence classes of 3-dimensional Lie algebras over the complex numbers was<br />

then refined by him to the real case by adding several types, giving Bianchi’s canonical<br />

form for the generating Lie algebra commutation relations for each type designated by<br />

consecutive Roman numerals I through IX, now known as the Bianchi classification.<br />

Below is Bianchi’s frist classification of all possible spaces which admit continous<br />

groups of motions. They are divided into six categories according to the type of their<br />

complete group G of motions. Spaces whose group of motions is a G 1 are class (A), classe<br />

(B) are spaces <strong>with</strong> group of motions G 2 , they are class (C) when it is an intransitive<br />

G 3 . The other two categories (D) , (E) contain the spaces whose group of motions is<br />

1 Original title: Sugli spazi a tre dimensioni che ammettono un gruppo continuo di movimenti,<br />

Memorie di Matematica e di Fisica della Societa Italiana delle Scienze, Serie Terza, Tomo XI, pp. 267–<br />

352 (1898). I found an english translation of this paper Translated by Robert Jantzen. This paper<br />

was also reprinted in: Opere [The Collected Works of Luigi Bianchi], Rome, Edizione Cremonese, 1952,<br />

vol. 9, pp. 17-109.<br />

158


E.0 159<br />

transitive, (D) those <strong>with</strong> a G 3 , and (E) those <strong>with</strong> a G 4 . The sixth category (F) includes<br />

the spaces of constant curvature which are maximally symmetric spaces which admit a<br />

group G 6 of motions. He proves the impossibility of other spaces <strong>with</strong> continuous groups<br />

of motions. For example that there does not exist any 3-dimensional space whose group<br />

of motions contains a real 5-parameter subgroup.<br />

Category A<br />

Groups G 1<br />

Type I<br />

ds 2 = Σ a ik dx i dx k<br />

<strong>with</strong> coefficients a ik independent of x 1<br />

group:<br />

X 1 f = ∂f<br />

∂x 1<br />

Or, the group G 3 is abelian and offers the first and simplest composition<br />

[X 1 , X 2 ]f = [X 1 , X 3 ]f = [X 2 , X 3 ]f = 0 .<br />

Category B<br />

Groups G 2<br />

ds 2 = dx 2 1 + α dx2 2 + 2β dx 2dx 3 + γ dx 2 3<br />

<strong>with</strong> α, β, γ functions only of x 1<br />

group:<br />

composition:<br />

X 1 f = ∂f<br />

∂x 2<br />

, X 2 f = ∂f<br />

∂x 3<br />

[X 1 , X 2 ] = 0<br />

ds 2 = dx 2 1 + α dx 2 2 + 2(β − αx 2 ) dx 2 dx 3 + (αx 2 2 − 2βx 2 + γ) dx 2 3


E.0 160<br />

<strong>with</strong> α, β, γ functions of x 1<br />

group:<br />

composition:<br />

X 1 f = ∂f<br />

∂x 3<br />

, X 2 f = e x 3 ∂f<br />

∂x 2<br />

[X 1 , X 2 ]f = X 2 f<br />

Category C<br />

Groups G 3 intransitive<br />

α) ds 2 = dx 2 1 + ϕ2 (x 1 )(dx 2 2 + dx2 3 )<br />

<strong>with</strong> ϕ(x 1 ) an arbitrary function of x 1<br />

group:<br />

composition:<br />

X 1 f = ∂f<br />

∂x 2<br />

, X 2 f = ∂f<br />

∂x 3<br />

, X 3 f = x 3<br />

∂f<br />

∂x 2<br />

− x 2<br />

∂f<br />

∂x 3<br />

[X 1 , X 2 ]f = 0 , [X 1 , X 3 ]f = −X 2 f , [X 2 , X 3 ]f = X 1 f<br />

β) ds 2 = dx 2 1 + ϕ2 (x 1 )(dx 2 2 + sin2 x 2 dx 3 ) 2<br />

group:<br />

composition:<br />

X 1 f = ∂f<br />

∂x 3<br />

, X 2 f = sin x 3<br />

∂f<br />

∂x 2<br />

+ cot x 2 cos x 3<br />

∂f<br />

∂x 3<br />

,<br />

X 3 f = cos x 3<br />

∂f<br />

∂x 2<br />

− cot x 2 sin x 3<br />

∂f<br />

∂x 3<br />

[X 1 , X 2 ]f = X 3 f , [X 2 , X 3 ]f = X 1 f , [X 3 , X 1 ]f = X 2 f<br />

γ) ds 2 = dx 2 1 + ϕ2 (x 1 ) (dx 2 2 + e2x 2<br />

dx 3 ) 2


E.0 161<br />

group:<br />

composition:<br />

X 1 f = ∂f<br />

∂x 3<br />

, X 2 f = ∂f<br />

∂x 2<br />

− x 3<br />

∂f<br />

∂x 3<br />

,<br />

X 3 f = x 3<br />

∂f<br />

∂x 2<br />

+ 1 2 (e−2x 2<br />

− x 2 3) ∂f<br />

∂x 3<br />

[X 1 , X 2 ]f = −X 1 f , [X 1 , X 3 ]f = X 2 f , [X 2 , X 3 ]f = −X 3 f<br />

Category D<br />

Groups G 3 transitive<br />

Type IV<br />

ds 2 = dx 2 1 + ex 1<br />

[dx 2 2 + 2x 1 dx 2 dx 3 + (x 2 1 + n2 ) dx 3 ) 2 ]<br />

group:<br />

composition:<br />

X 1 f = 2 ∂f<br />

∂x 2<br />

, X 2 f = ∂f<br />

∂x 3<br />

,<br />

X 3 f = −2 ∂f<br />

∂x 1<br />

+ (x 2 + 2x 3 ) ∂f<br />

∂x 2<br />

+ x 3<br />

∂f<br />

∂x 3<br />

[X 1 , X 2 ] = 0 , [X 1 , X 3 ]f = X 1 f , [X 2 , X 3 ]f = X 1 f + X 2 f<br />

Type VI<br />

ds 2 = dx 2 1 + e 2x 1<br />

dx 2 2 + 2ne (h+1)x 1<br />

dx 2 dx 3 + e 2hx 1<br />

dx 2 3<br />

group:<br />

composition:<br />

X 1 f = ∂f<br />

∂x 2<br />

, X 2 f = ∂f<br />

∂x 3<br />

,<br />

X 3 f = − ∂f<br />

∂x 1<br />

+ x 2<br />

∂f<br />

∂x 2<br />

+ hx 3<br />

∂f<br />

∂x 3<br />

[X 1 , X 2 ] = 0 , [X 1 , X 3 ]f = X 1 f , [X 2 , X 3 ]f = hX 2 f


E.0 162<br />

Type V II 1<br />

ds 2 = dx 2 1 + (n + cos x 1) dx 2 2 + 2 sin x 1 dx 2 dx 3 + (n − cos x 1 ) dx 2 3<br />

group:<br />

<strong>with</strong> the composition<br />

X 1 f = ∂f<br />

∂x 2<br />

, X 2 f = ∂f<br />

∂x 3<br />

, X 3 f = 2 ∂f<br />

∂x 1<br />

− x 3<br />

∂f<br />

∂x 2<br />

+ x 2<br />

∂f<br />

∂x 3<br />

,<br />

[X 1 , X 2 ]f = 0 , [X 1 , X 3 ]f = X 2 f , [X 2 , X 3 ]f = −X 1 f .<br />

Type V II 2<br />

ds 2 = dx 2 1 + e−hx 1<br />

{(n + cos vx 1 ) dx 2 2 + (h cos vx 1 + v sin x 1 + hn) dx 2 dx 3<br />

(<br />

) }<br />

2−v<br />

+<br />

2<br />

cos vx<br />

2 1 + hv sin vx 2 1 + n dx 2 3 , h ≠ 0 (0 < h < 2)<br />

where v = √ 4 − h 2<br />

group:<br />

composition:<br />

X 1 f = ∂f<br />

∂x 2<br />

, X 2 f = ∂f<br />

∂x 3<br />

,<br />

X 3 f = ∂f<br />

∂x 1<br />

− x 3<br />

∂f<br />

∂x 2<br />

+ (x 2 + hx 3 ) ∂f<br />

∂x 3<br />

[X 1 , X 2 ] = 0 , [X 1 , X 3 ]f = X 2 f , [X 2 , X 3 ]f = −X 1 f + hX 2 f<br />

Type VIII<br />

(<br />

)<br />

ds 2 = Q(4) (x 1 )<br />

dx 2 24 1 + Q(x 1 ) dx 2 2 + Q(x 1 )x 2 2 − Q′ (x 1 )<br />

x<br />

2 2 + Q′′ (x 1 )<br />

− h 2 2<br />

( )<br />

{ ( ) }<br />

+2 Q ′′ (x 1 )<br />

+ h dx<br />

12 1 dx 2 + 2 Q ′′′ (x 1 )<br />

− Q ′′ (x 1 )<br />

+ h x<br />

24 12 2 dx 1 dx 3<br />

( )<br />

+2 Q ′ (x 1 )<br />

− Q(x<br />

4 1 )x 2 dx 2 dx 3 ,<br />

<strong>with</strong> Q(x 1 ) a fourth degree polynomial in x 1 <strong>with</strong> its first<br />

coefficient positive (or zero), and h a constant<br />

group:<br />

X 1 f = e −x 3 ∂f<br />

∂x 1<br />

− x 2 2 e−x 3 ∂f<br />

∂x 2<br />

− 2x 2 e −x 3 ∂f<br />

∂x 3<br />

;<br />

X 2 f = ∂f<br />

∂x 3<br />

, X 3 f = ∂f<br />

∂x 2<br />

dx 2 3


E.0 163<br />

composition:<br />

[X 1 , X 2 ]f = X 1 f , [X 1 , X 3 ]f = 2X 2 f , [X 2 , X 3 ]f = X 3 f<br />

ds 2 = Σ i,k a ik dx i dx k<br />

group:<br />

composition:<br />

a 11 = 2e cos 2x 3 + 2f sin 2x 3 + (a 2 + d 2 )/2 ,<br />

a 22 = 2 sin x 1 cos x 1 (b sin x 3 − c cos x 3 ) − a 11 sin 2 x 1 + a 2 + d sin 2 x 1 ,<br />

a 33 = a 2 , a 13 = b cos x 3 + c sin x 3 ,<br />

a 12 = cos x 1 (b cos x 3 + c sin x 3 ) + 2 sin x 1 (e sin 2x 3 − f cos 2x 3 ) ,<br />

a 23 = a 2 cos x 1 + sin x 1 (b sin x 3 − c cos x 3 )<br />

X 1 f = ∂f<br />

∂x 2<br />

, X 2 f = cos x 2<br />

∂f<br />

∂x 1<br />

− cot x 1 sin x 2<br />

∂f<br />

∂x 2<br />

+ sin x 2<br />

sin x 1<br />

∂f<br />

∂x 3<br />

,<br />

X 3 f = − sin x 2<br />

∂f<br />

∂x 1<br />

− cot x 1 cos x 2<br />

∂f<br />

∂x 2<br />

+ cos x 2<br />

sin x 1<br />

∂f<br />

∂x 3<br />

[X 1 , X 2 ]f = X 3 f , [X 2 , X 3 ]f = X 1 f , [X 3 , X 1 ]f = X 2 f<br />

Category E<br />

Groups G 4<br />

a) Type II<br />

ds 2 = dx 2 1 + dx 2 2 + 2x 1 dx 2 dx 3 + (x 2 1 + 1) dx 2 3<br />

group:<br />

composition:<br />

X 1 f = ∂f<br />

∂x 2<br />

, X 2 f = ∂f<br />

∂x 3<br />

, X 3 f = − ∂f<br />

∂x 1<br />

+ x 3<br />

∂f<br />

∂x 2<br />

,<br />

X 4 f = x 3<br />

∂f<br />

∂x 1<br />

+ 1 2 (x2 1 − x2 3 ) ∂f<br />

∂x 2<br />

− x 1<br />

∂f<br />

∂x 3<br />

[X 1 , X 2 ] = [X 1 , X 3 ] = [X 1 , X 4 ] = 0 ,<br />

[X 2 , X 3 ]f = X 1 f , [X 2 , X 4 ]f = −X 3 f , [X 3 , X 4 ]f = X 2 f


E.0 164<br />

b) Types III, VIII<br />

ds 2 = dx 2 1 + e2x 1<br />

dx 2 2 + 2nex 1<br />

dx 2 dx 3 + dx 2 3<br />

group:<br />

composition:<br />

X 1 f = ∂f , X 2 f = ∂f , X 3 f = ∂f ∂f<br />

− x 2 ,<br />

∂x 2 ∂x 3 ∂x 1 ∂x 2<br />

∂f<br />

X 4 f = x 2 + 1 ( e<br />

−2x 1<br />

) ∂f<br />

∂x 1 2 1 − n − 2 x2 2 − ne−x 1<br />

∂f<br />

∂x 2 1 − n 2 ∂x 3<br />

[X 1 , X 2 ] = 0 , [X 1 , X 3 ]f = −X 1 f , [X 1 , X 4 ]f = X 3 f ,<br />

[X 2 , X 3 ] = 0 , [X 2 , X 4 ] = 0 , [X 3 , X 4 ]f = −X 4 f<br />

c) Type IX<br />

ds 2 = dx 2 1 + (sin2 x 1 + n 2 cos 2 x 1 ) dx 2 2 + 2n cos x 1 dx 2 dx 3 + dx 2 3<br />

group:<br />

composition:<br />

X 1 f = ∂f<br />

∂f<br />

∂f<br />

, X 2 f = cos x 2 − cot x 1 sin x 2 + n sin x 2<br />

∂x 2 ∂x 1 ∂x 2 sin x 1<br />

X 3 f = − sin x 2<br />

∂f<br />

∂x 1<br />

− cot x 1 cos x 2<br />

∂f<br />

∂x 2<br />

+ n cos x 2<br />

sin x 1<br />

∂f<br />

∂x 3<br />

,<br />

∂f<br />

, X 4 f = ∂f<br />

∂x 3 ∂x 3<br />

[X 1 , X 2 ]f = X 3 f , [X 2 , X 3 ]f = X 1 f , [X 3 , X 1 ]f = X 2 f ,<br />

[X 1 , X 4 ] = [X 2 , X 4 ] = [X 3 , X 4 ] = 0<br />

Category F<br />

Groups G 6 — spaces of constant curvature


AppendixF<br />

Simulating <strong>CMB</strong> maps for non-SI<br />

correlation functions<br />

We define a zero-mean Gaussian random field T over a grid S <strong>with</strong> covariance function<br />

C : R 2 → R<br />

(F.1)<br />

Having such a field means that for any finite set of n points<br />

x 1 , x 2 , · · · , x n ∈ S<br />

(F.2)<br />

we have random variables<br />

T (x 1 ), T (x 2 ), · · · , T (x n );<br />

(F.3)<br />

which are jointly normal <strong>with</strong> mean zero and covariances<br />

Cov(T (x i ), T (x j )) = C(T (x i ), T (x j )).<br />

(F.4)<br />

Lets denote T (x i ) by T i for simplicity and also the covariance matrix C(T (x i ), T (x j ))<br />

by C ij . The distribution function for this field is given by<br />

P =<br />

1<br />

(2π) n/2√ ‖C‖ exp (−1 2 T iC −1<br />

ij T j).<br />

(F.5)<br />

In the above expression we have made use of Einstein summation convention to drop<br />

the summation sign, hereafter we will stick to this notation. And by ‖C‖ we mean the<br />

determinant of C ij .<br />

165


F.0 166<br />

In eqn. (F.5) we can not write the distribution function as a product of n Gaussian<br />

distributions because the covariance matrix in general may not be diagonal and it will<br />

have cross correlation terms.<br />

There are two ways to proceed further. The diagonalization method and the Cholesky<br />

decomposition.<br />

Diagonalization method<br />

The first way is the following: suppose we can diagonalize C −1 which means we<br />

can find a unitary matrix U such that U −1 C −1 U is diagonal. In that case we can<br />

immediately write the exponent in (F.5) as a sum of independent quantities and the<br />

distribution function will break into a product of n Gaussian distributions.<br />

If we use some matrix D to change C ij into diagonal form say, diag(σ 1 , σ 2 , ..., σ N )<br />

then the power of e would change as:<br />

T T<br />

i U T UC −1<br />

ij U T UT i = ψ 2 i /σ2 i<br />

(F.6)<br />

and ψ i are nothing but elements of rotated T vector under the effect of U ( −→ ψ = U −→ T ).<br />

So P (T i ) can be written as:<br />

P (T i ) = Π i<br />

1<br />

(2πσ i ) 1/2 e−ψ2 i /σ2 i<br />

P i (ψ i ) =<br />

= Π i P i (ψ i ) ,<br />

1<br />

√<br />

2πσi<br />

e − ψ2 i<br />

2σ 2 i<br />

(F.7)<br />

So now everything is clear: it is enough to generate a vector −→ ψ whose elements are<br />

gaussian random numbers, then multiply each element by the corresponding eigenvalue<br />

of C ij say σ i to find a gaussian variable <strong>with</strong> variance σ i . Then apply an inverse rotation<br />

on vector −→ ψ to find T = U −1−→ ψ which is a realization of the temperature fluctuations<br />

on the sky.<br />

This has been done using the method of Householder reduction of a real, symmetric,<br />

N × N matrix into a tridiagonal matrix. And then the method of QL to diagonalize<br />

that tridiagonal matrix. An example of the use of this method is simulation of noise


F.0 167<br />

maps using the noise covariance matrix given in Fig. 6.1 which is ofcourse diagonal by<br />

itself and hence is a trivial example! The resultant map is shown in Fig. F.1.


F.0 168<br />

Figure F.1 A simulated map from noise covariance matrix using inversion method.


F.0 169<br />

Cholesky decomposition<br />

There is another way of proceeding further <strong>with</strong> the equation(F.5). In matrix notation<br />

we can write it as<br />

P (T ) =<br />

1<br />

(2π) n/2√ ‖C‖ exp (−1 2 T t C −1 T )<br />

(F.8)<br />

where T t means T transposed. Now if there exists a matrix L such C can be decomposed<br />

into L and L t in the following manner<br />

C = L L t .<br />

(F.9)<br />

Then we can substitute this for C in the exponent of (F.8) to get<br />

T t C −1 T = T t (LL t ) −1 T (F.10)<br />

= (L −1 T ) t L −1 T<br />

= Y t Y<br />

Where<br />

Y = L −1 T or T = L Y. (F.11)<br />

The distribution function can be written for Y<br />

P (Y ) =<br />

=<br />

=<br />

1<br />

(2π) n/2√ ‖C‖ exp (−1 2 Y t Y ) ‖∂T/∂Y ‖<br />

1<br />

(2π) exp (−1 ∑<br />

y n/2 i 2 2<br />

)<br />

i<br />

n∏<br />

{ 1<br />

(2π) exp (−1 2 y2 i )}<br />

i=1<br />

(F.12)<br />

This is nothing but the product of n independent Gaussian distributions for a single<br />

variable. Based on the above procedure we can prescribe a way to make realizations of<br />

a Gaussian random field <strong>with</strong> a given covariance matrix:<br />

Recipe: Map Making<br />

• Decompose the covariance matrix into L and L t . In mathematics this<br />

is called the Cholesky decomposition of a positive definite matrix. In<br />

Cholesky decomposition one decomposes a positive definite matrix into


F.0 170<br />

a triangular matrix and its transposed. There exists a subroutine<br />

choldc.f in Numerical Recipes which takes the matrix as an input and<br />

in output it returns the triangular matrix and in the diagonal it returns<br />

the square root of the eigenvalues of the matrix.<br />

• Generate n independent Gaussian random variables(Y ).<br />

• Use L to rotate back the vector Y and get the joint Gaussian random<br />

variables T = LY .<br />

• The computational effort in the one-time Cholesky decomposition scales<br />

as n 3 . However, Cholesky decomposition is significantly faster than the<br />

other alternative of diagonalization.<br />

This method has some advantages to the previous methods of generating <strong>CMB</strong> sky<br />

maps which use the angular power spectrum C l to generate <strong>CMB</strong> sky maps. Below we<br />

list a number of such advantages:<br />

1. The first advantage is that in this method we are using the covariance matrix<br />

to generate the <strong>CMB</strong> map. This is very important when we deal <strong>with</strong> statistically<br />

anisotropic models such as models <strong>with</strong> non-trivial topologies, or when we<br />

want to account effects of non-circular beam, foreground residuals and statistically<br />

anisotropic noise. In these models we can not use the angular power spectrum because<br />

we don’t have statistical isotropy to let us expand the covariance in terms<br />

of power spectrum. In the statistically non-isotropic models, power spectrum does<br />

not carry all the information of the random field and can not be used to generate<br />

a valid realization of the random field.<br />

2. The other advantage in this method is that we can generate a map for a small<br />

part of the sky <strong>with</strong> the high resolution at a proportional small fraction of the<br />

cost of generating the whole sky map. It is just enough to have the covariance<br />

matrix of the region of our interest, and making map <strong>with</strong>out generating the full<br />

sky map is as we described in the recipe. This is in stark contrast to generating<br />

maps using the power spectrum. In that case the computational expense is high<br />

for high resolution even if we are simulating a small fraction of the sky.


F.0 171<br />

Figure F.2 A simulated <strong>CMB</strong> map in a toroidal universe using Cholesky decomposition<br />

method. Only naive Sachs-Wolfe effect has been considered and R SLS = 0.65L.


AppendixG<br />

Useful Standard Integrals and Summation<br />

Rules<br />

For completeness we list a set of standard integrals and summation rules which are<br />

needed in BiPS calculations. Orthonormality of spherical harmonics<br />

∫<br />

∫<br />

dΩˆn Y lm (ˆn) = √ 4πδ l0 δ m0 , (G.1)<br />

dΩˆn Y l1 m 1<br />

(ˆn)Y ∗<br />

l 2 m 2<br />

(ˆn) = δ l1 l 2<br />

δ m1 m 2<br />

, (G.2)<br />

∫<br />

dΩˆn Y l1 m 1<br />

(ˆn)Y l2 m 2<br />

(ˆn)Y ∗<br />

l 3 m 3<br />

(ˆn) =<br />

√<br />

(2l 1 + 1)(2l 2 + 1)<br />

C l 30<br />

l<br />

4π(2l 3 + 1) 1 0l 2 0 Cl 3m 3<br />

l 1 m 1 l 2 m 2<br />

.<br />

(G.3)<br />

(G.4)<br />

Symmetry property of spherical harmonics<br />

Y ∗<br />

lm(ˆn) = (−1) m Y l−m (ˆn).<br />

(G.5)<br />

Spherical harmonic expansion of Legendre polynomials<br />

P l (ˆn · ˆn ′ ) =<br />

4π<br />

2l + 1<br />

l∑<br />

m=−l<br />

Y ∗<br />

lm (ˆn)Y lm(ˆn ′ ).<br />

(G.6)<br />

Orthonormality relation of χ l (ω)<br />

∫ π<br />

0<br />

χ l 1<br />

(ω)χ l 2<br />

(ω) sin 2 ω 2 dω = π 2 δ l 1 l 2<br />

.<br />

(G.7)<br />

172


G.0 173<br />

Orthonormality of D l MM (R)<br />

∫<br />

dRD ∗l<br />

MM ′(R)Dl mm ′(R) =<br />

∑<br />

8π2<br />

2l + 1 δ llδ mM δ m ′ M ′,<br />

mm ′ D l mm ′(R 1)D l m ′ m (R 2) = χ l (R 1 R 2 )<br />

Symmetry properties of Clebsch-Gordan coefficients<br />

C cγ<br />

aαbβ<br />

= (−1) a+b−c C cγ<br />

bβaα ,<br />

C cγ<br />

aαbβ<br />

= (−1) a+b−c C c−γ<br />

a−αb−β .<br />

(G.8)<br />

(G.9)<br />

(G.10)<br />

Summation rules of Clebsch-Gordan coefficients<br />

∑<br />

αβ<br />

∑<br />

aγ<br />

∑<br />

cγ<br />

C cγ<br />

aαbβ Cc′ γ ′<br />

aαbβ<br />

= δ cc ′δ γγ ′{abc}{abc ′ }<br />

C cγ<br />

aαbβ Ccγ aαb ′ β ′ = 2c + 1<br />

2b + 1 δ bb ′δ ββ ′{abc}{ab′ c}<br />

C cγ<br />

aαbβ Ccγ aα ′ bβ ′ = δ αα ′δ ββ ′{abc}, (G.11)<br />

where the 3j-symbol {abc} is defined by<br />

⎧<br />

⎨ 1 if a + b + c is integer and |a − b| ≤ c ≤ (a + b),<br />

{abc} =<br />

⎩ 0 otherwise,


Bibliography<br />

R. Aurich, S. Lustig and F. Steiner, Class. Quant. Grav. 22, 2061 (2005) [arXiv:astroph/0412569].<br />

K. Bamba and J. Yokoyama, Phys.Rev. D69 (2004) 043507.<br />

J. M. Bardeen, P. J. Steinhardt, & M. S. Turner, 1983, Phys. Rev. D 28, 679.<br />

J. D. Barrow, R. Juszkiewicz and D. H. Sonoda, 1985, Mon. Not. R. astr. Soc., 213,<br />

917.<br />

N. Bartolo, E. Komatsu, S. Matarrese and A. Riotto, preprint (astro-ph/0406398)<br />

C. L. Bennett, et.al., 2003, Astrophys. J. Suppl., 148, 1.<br />

A. Bernui, B. Mota, M. J. Reboucas and R. Tavakol, arXiv:astro-ph/0511666.<br />

J. R. Bond, et al. 2004, Int. J. Theor. Phys. 2004, ed. E. Verdaguer, ”The Early Universe:<br />

Confronting theory <strong>with</strong> observations” (June 21-27, 2003) (astro-ph/0406195).<br />

J. R. Bond, D. Pogosyan, & T. Souradeep, 1998, Class. Quant. Grav. 15, 2671; ibid.<br />

2000, Phys. Rev. D 62,043005.<br />

J. R. Bond, D. Pogosyan, & T. Souradeep, 2000, Phys. Rev. D 62, 043006<br />

R. Bowen & P. G. Ferreira, 2002, Phys. Rev. D 66, 041302.<br />

S. L. Bridle, O. Lahav, J. P. Ostriker and P. J. Steinhardt, Science 299, 1532 (2003)<br />

[arXiv:astro-ph/0303180].<br />

E. F. Bunn, P. Ferreira and J. Silk, 1996, Phys.Rev.Lett. 77, 2883<br />

G. Chen, P. Mukherjee, T. Kahniashvili, B. Ratra and Y. Wang, Astrophys. J. 611<br />

(2004) 655.<br />

P. Coles, et al. 2003, preprint (astro-ph/0310252).<br />

C. B. Collins and S. W. Hawking, 1973, Astrophys. J. 180, 317.<br />

C. B. Collins and S. W. Hawking, 1973, Mon. Not. R. astr. Soc. 162, 307.<br />

C. J. Copi, D. Huterer, & G. D. Starkman, 2004, Phys. Rev. D. in press, (astroph/0310511).<br />

N. J. Cornish, D. N. Spergel, & G. D. Starkman, 1998, Class. Quantum Grav., 15, 2657<br />

N. J. Cornish, D. Spergel, G. Starkman and E. Komatsu, 2004, Phys. Rev. Lett. 92,<br />

174


G.0 BIBLIOGRAPHY 175<br />

201302.<br />

M. Cruz, E. Martinez-Gonzalez, P. Vielva and L. Cayon, preprint (astro-ph/0405341).<br />

A. de Oliviera Costa and G. F. Smoot, Astrophys. J. 448, 477 (1995)<br />

A. de Oliveira-Costa, G. F. Smoot and A. A. Starobinsky, 1996, ApJ 468, 457.<br />

A. de Oliveira-Costa, M. Tegmark, M. Zaldarriaga, & A. Hamilton, 2004, Phys.<br />

Rev.D69, 063516.<br />

P. Dineen, G. Rocha and P, Coles, preprint (astro-ph/0404356).<br />

E. P. Donoghue and J. F. Donoghue, preprint (astro-ph/0411237).<br />

A. G. Doroshkevich, V. N. Lukash and I. D. Novikov, 1975, Soviet Astr., 18, 554.<br />

R. Durrer, T. Kahniashvili and A. Yates, 1998, Phys. Rev. D 58, 3004.<br />

G. F. R. Ellis, 1971, Gen. Rel. Grav. 2, 7.<br />

G. F. R. Ellis and M. A. H. MacCallum, 1969, Commun. math Phys., 12, 108.<br />

G. F. R. Ellis & H. van Elst, 1999, NATO ASIC Proc. 541: Theoretical and Observational<br />

<strong>Cosmology</strong> , 1<br />

H. K. Eriksen, et al., 2004, Astrophys. J 605, 14.<br />

H. K. Eriksen, et al., 2004, Astrophys. J. 612, 64.<br />

H. K. Eriksen, et al., 2004, Astrophys. J. 612, 633.<br />

P. Ferreira & J. Magueijo, 1997, Phys. Rev. D56 4578.<br />

E. Gausmann, R. Lehoucq, J. P. Luminet, J. P. Uzan and J. Weeks, Class. Quant. Grav.<br />

18, 5155 (2001)<br />

E. Gaztanaga & J. Wagg, 2003, Phys. Rev. D68 021302.<br />

G. I. Gomero, A. F. F. Teixeira, M. J. Reboucas and A. Bernui, Int. J. Mod. Phys. D 11,<br />

869 (2002) [arXiv:gr-qc/9811038], G. I. Gomero, M. J. Reboucas and A. F. F. Teixeira,<br />

Phys. Lett. A 275, 355 (2000) [arXiv:gr-qc/9909078], G. I. Gomero, M. J. Reboucas<br />

and A. F. F. Teixeira, Class. Quant. Grav. 18, 1885 (2001) [arXiv:gr-qc/9911049],<br />

G. I. Gomero, M. J. Reboucas and R. K. Tavakol, Class. Quant. Grav. 18, 4461<br />

(2001) [arXiv:gr-qc/0105002], G. I. Gomero, M. J. Reboucas and A. F. F. Teixeira,<br />

Int. J. Mod. Phys. D 9, 687 (2000) [arXiv:gr-qc/0105048].<br />

G. I. Gomero and M. J. Reboucas, Phys. Lett. A 311, 319 (2003) [arXiv:gr-qc/0202094],<br />

G. I. Gomero, M. J. Reboucas and R. Tavakol, Int. J. Mod. Phys. A 17, 4261 (2002)<br />

[arXiv:gr-qc/0210016].<br />

K. M. Górski, E. Hivon and B. D. Wandelt, 1999, in ”Evolution of Large-Scale Structure”,<br />

eds. A. J. Banday, R. S. Sheth and L. Da Costa, PrintPartners Ipskamp, NL,<br />

pp. 37-42 (also astro-ph/9812350).<br />

M. Goliath and G. F. R. Ellis, Phys. Rev. D 60, 023502 (1999) [arXiv:gr-qc/9811068].<br />

T. Ghosh, T. Souradeep and A. Hajian, 2006, in preparation.<br />

A. H. Guth & S. Y. Pi, 1982, Phys. Rev. Lett., 49, 1110.<br />

A. Hajian & T. Souradeep, 2003a preprint (astro-ph/0301590).


G.0 BIBLIOGRAPHY 176<br />

A. Hajian & T. Souradeep, 2003b, ApJ 597, L5 (2003).<br />

A. Hajian & T. Souradeep, 2005, ApJ 618, L63.<br />

A. Hajian, D. Pogosyan, T. Souradeep, C. Contaldi and R. Bond, 2004 in preparation;<br />

Proc. 20th IAP Colloquium on Cosmic Microwave Background physics and observation,<br />

2004.<br />

A. Hajian & T. Souradeep, 2005, preprint. (astro-ph/0501001).<br />

F. K. Hansen, A. J. Banday & K. M. Gorski, 2004, preprint. (astro-ph/0404206)<br />

S. W. Hawking, Mon. Not. Roy. Astron. Soc. 142, 129 (1969).<br />

G. Hinshaw, et al., Astrophys. J. Suppl.,(2003) ,148, 135.<br />

W. S. Hipolito-Ricaldi and G. I. Gomero, Phys. Rev. D 72, 103008 (2005) [arXiv:astroph/0507238].<br />

T. R. Jaffe, et al., 2005, preprint, atro-ph/0503213.<br />

E. Komatsu, et.al., 2003, ApJS, 148, 119.<br />

A. Kogut, et al., 2003, Astrophys. J. Suppl.,148, 161.<br />

A. Kogut, G. Hinshaw and A. J. Banday, 1997, Phys.Rev. D55, 1901<br />

E. Kourkchi, Masters Thesis, Sharif University of Technology, 2006.<br />

M. Lachieze-Rey and J. P. Luminet, 1995, Phys. Rep. 25, 136.<br />

M. Lachieze-Rey, Class. Quant. Grav. 21, 2455 (2004) [arXiv:gr-qc/0402035].<br />

D. L. Larson & B. D. Wandelt, 2004, preprint, (astro-ph/0404037).<br />

J. Levin, 2002, Phys. Rep. 365, 251.<br />

J. Levin, E. Scannapieco and J. Silk, (1998), Class.Quant.Grav. 15, 2689.<br />

A. Linde, (2004), JCAP 0410, 004, (astro-ph/0408164).<br />

J. P. Luminet, et al., 2003, Nature 425 593.<br />

J. P. Luminet, arXiv:astro-ph/0501189.<br />

C. P. Ma & E. Bertschinger, 1995, ApJ, 455, 7<br />

J. D. McEwen, M. P. Hobson, A. N. Lasenby and D. J. Mortlock, Mon. Not. Roy. Astron.<br />

Soc. 359, 1583 (2005) [arXiv:astro-ph/0406604].<br />

J. D. McEwen, M. P. Hobson, A. N. Lasenby and D. J. Mortlock, arXiv:astroph/0510349.<br />

C. W. Misner, 1968, ApJ, 151, 431<br />

S. Mitra, A. S. Sengupta and T. Souradeep, Phys. Rev. D 70, 103002 (2004).<br />

B. Mota, G. I. Gomero, M. J. Reboucas and R. Tavakol, Class. Quant. Grav. 21, 3361<br />

(2004) [arXiv:astro-ph/0309371].<br />

H. Mukhopadhyay, IUCAA graduate school project report, available at<br />

http://meghnad.<strong>iucaa</strong>.ernet.in/ tarun/papers/students/himan project.pdf<br />

D. Munshi, T. Souradeep and A. A. Starobinsky, (1995), Astrophys. J., 454, 552.<br />

M. Nakahara, 1990, Bristol : Adam and Hilger, 1990.<br />

Naselsky, et al., 2004, preprint, (astro-ph/0405523); ibid (astro-ph/0405181), 2003;<br />

ApJ,599,L53 and references therein.<br />

L. Page, et al., 2003, Astrophys. J. Suppl.,148, 233.


G.0 BIBLIOGRAPHY 177<br />

C. Park, 2004, Mon. Not. Roy. Astron. Soc. 349, 313.<br />

H. V. Peiris, et al., 2003, Astrophys. J. Suppl.,148, 213.<br />

S. Prunet, J. Uzan, F. Bernardeau and T. Brunier, 2004, preprint (astro-ph/0406364).<br />

B. Ratra, (1992), Astrophys.J. 391, L1.<br />

A. Riazuelo, J. P. Uzan, R. Lehoucq and J. Weeks, Phys. Rev. D 69, 103514 (2004)<br />

A. Riazuelo, J. Weeks, J. P. Uzan, R. Lehoucq and J. P. Luminet, Phys. Rev. D 69,<br />

103518 (2004)<br />

B. F. Roukema, B. Lew, M. Cechowska, A. Marecki and S. Bajtlik, Astron. Astrophys.<br />

423, 821 (2004) [arXiv:astro-ph/0402608].<br />

R. K. Sachs and A. M. Wolfe, 1967, ApJ, 147, 73.<br />

R. Saha, A. Hajian, T. Souradeep and P. Jain, 2006, in preparation.<br />

D. Scott, arXiv:astro-ph/0510731.<br />

D. J. Schwarz, et al., 2004, preprint (astro-ph/0403353).<br />

A. Shafieloo & T. Souradeep, 2004, Phys. Rev. D 70, 043523 (2004).<br />

T. Souradeep, 2000, in ‘The Universe’, eds. Dadhich, N. & Kembhavi, A., Kluwer.<br />

T. Souradeep and A. Hajian, 2004, Pramana, 62, 793.<br />

D. Spergel, et al., 2003, Astrophys. J. Suppl., 148, 175.<br />

D. Spergel and D. M. Goldberg, (1999), Phys.Rev. D59, 103001.<br />

G. Starkman, Class. 1998, Quantum Grav. 15, 2529.<br />

A. A. Starobinsky, 1982, Phys. Lett, 117B, 175.<br />

A. A. Starobinsky, JETP Lett. 57, 622 (1993)<br />

M. Tegmark, A. de Oliveira-Costa & A. Hamilton, 2004, Phys.Rev. D68 123523.<br />

D. A. Varshalovich, A. N. Moskalev and V. K. Khersonskii, 1988 Quantum Theory of<br />

Angular Momentum (World Scientific).<br />

P. Vielva, E. Martinez-Gonzalez, R. B. Barreiro, J. L. Sanz and L. Cayon, 2003, preprint<br />

(astro-ph/0310273).<br />

R. M. Wald, 1983, Phys. Rev. D, 28, 2118.<br />

R. M. Wald, 1984, Chicago: University of Chicago Press, 1984.<br />

J. R. Weeks, Mod. Phys. Lett. A 18, 2099 (2003) [arXiv:astro-ph/0212006].<br />

J. Weeks, R. Lehoucq and J. P. Uzan, Class. Quant. Grav. 20, 1529 (2003)<br />

J. Weeks, J. P. Luminet, A. Riazuelo and R. Lehoucq, Mon. Not. Roy. Astron. Soc. 352,<br />

258 (2004) [arXiv:astro-ph/0312312].<br />

J. R. Weeks, arXiv:math.SP/0502566.<br />

M. Zaldarriaga, U. Seljak and E. Bertschinger, 1998, Astrophys. J., 494, 491.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!