Statistics of The Galaxy Distribution
Statistics of The Galaxy Distribution
Statistics of The Galaxy Distribution
of the GALAXY
DISTRIBUTION
STATISTICS
of the GALAXY
DISTRIBUTION
Vicent J. Martnez
Enn Saar
2001028885
This book contains information obtained from authentic and highly regarded sources. Reprinted material
is quoted with permission, and sources are indicated. A wide variety of references are listed. Reasonable
efforts have been made to publish reliable data and information, but the author and the publisher cannot
assume responsibility for the validity of all materials or for the consequences of their use.
Neither this book nor any part may be reproduced or transmitted in any form or by any means, electronic
or mechanical, including photocopying, microfilming, and recording, or by any information storage or
retrieval system, without prior permission in writing from the publisher.
The consent of CRC Press LLC does not extend to copying for general distribution, for promotion, for
creating new works, or for resale. Specific permission must be obtained in writing from CRC Press LLC
for such copying.
Direct all inquiries to CRC Press LLC, 2000 N.W. Corporate Blvd., Boca Raton, Florida 33431.
Trademark Notice: Product or corporate names may be trademarks or registered trademarks, and are
used only for identification and explanation, without intent to infringe.
Contents
Preface
Acknowledgments
1 The clumpy universe
1.1 Galaxies
1.1.1 The Milky Way Galaxy
1.1.2 Morphological classification and properties
1.1.3 Brightness and magnitude systems
1.1.4 Distance estimators
1.2 Mapping the universe
1.2.1 Redshift surveys
1.2.2 Peculiar motions
1.3 Selection effects and biases
1.3.1 Galaxy obscuration
1.3.2 Flux limit, luminosity function, and selection function
1.3.3 Segregation
1.3.4 Cosmic variance
1.3.5 Malmquist bias
1.3.6 K-correction
1.3.7 Velocity corrections
1.4 Current and future galaxy catalogs
1.4.1 The galaxy distribution in projection
1.4.2 The three-dimensional galaxy distribution
1.5 The observed structures: clusters, filaments, walls, and voids
1.5.1 Groups and clusters of galaxies
1.5.2 Catalogs of clusters of galaxies
1.5.3 Superclusters: filaments and walls of galaxies
1.5.4 Voids
1.5.5 The texture of the galaxy distribution
2 The standard model of the universe
2.1 Introduction
7.4
7.5
7.6
7.7
7.8
7.9
8.7.3
8.7.4
8.7.5
8.7.6
Power spectra
Lucy deconvolution
Wide-angle surveys
Pencil-beams and slices
9 Cosmography
9.1 Introduction
9.2 Potent method
9.3 Wiener filtering
9.3.1 Filtering in spherical basis
9.3.2 Density interpolation
9.3.3 Wiener reconstruction
9.3.4 Maps
9.3.5 Velocity reconstruction
9.4 Constrained fields
9.4.1 Constrained realizations for models
9.5 Time machines
9.6 Gravitational lensing
9.6.1 Physics of gravitational lensing
9.6.2 Weak lensing
9.6.3 Cosmic shear
10 Structure statistics
10.1 Introduction
10.2 Topological description
10.2.1 The theory of topological analysis: the genus
10.2.2 Estimation of the topology, technicalities
10.2.3 Topological measurements: observations
10.3 Structure functions
10.3.1 Three-dimensional shape statistics
10.3.2 Minkowski functionals
10.4 Cluster and percolation analysis
10.5 Minimal spanning trees
10.6 Wavelets
10.6.1 Wavelet theory
10.6.2 Wavelets and multifractals
10.7 Cluster-finding algorithms
10.7.1 MST
10.7.2 Modified friends-of-friends
10.7.3 Wavelets
10.8 Void statistics
10.9 Checking for periodicity
10
Preface
You may ask What can a hard headed statistician offer to a starry eyed astronomer?
The answer is, Plenty. One normally associates statistics with large numbers, and
astronomy is full of large numbers. The number of stars in our galaxy, the so-called
Milky Way System, is more than a hundred thousand million. The number of galaxies
in the observable universe is upwards of a thousand million. Surely these large numbers
justify a prima-facie case for the use of statistical techniques! . . . I have every reason to
believe that increased interaction between statistics and astronomy will be to the benefit
of both the subjects.
J. V. Narlikar
The last half of the twentieth century saw cosmology develop into a very active
and diverse field of science. This was largely due to the development of observational techniques that allowed astronomers to observe extremely distant regions
of space. This motivated a flow of new theories about the evolution of our universe and the formation of the large-scale structure we found in it. The main tools
to compare theoretical results with observations in astronomy are statistical, so
the new theories and observations also initiated an active use of spatial statistics
in cosmology.
Many of the statistical methods used in the analysis of the large-scale distribution of matter in the universe have been developed by cosmologists and are not too
rigorous. In many cases, similar methods, sometimes under different names, had
been used for years in mainstream spatial statistics. In the late 1950s, when the
Berkeley statisticians J. Neyman and E. Scott carried out an intensive program for
the analysis of galaxy catalogs, the connection between spatial statisticians and
cosmologists was a fruitful one. However, in the following 30 years cosmologists
were not, in general, aware of developments in statistics, and vice versa. Fortunately, recent years have brought the resumption of a dialog between astronomers
and mathematicians, led by the Penn State conferences. We hope that this dialog will continue and will be useful. Cosmology is a good field for applications
of spatial statistics. Its well-defined and growing data sets represent an important
challenge for the statistical analysis, and therefore for the mathematical community.
11
The very influential book by Peebles (1980), a milestone in the field, gave the
first complete description of statistical methods for study of the spatial distribution of galaxies and of the essential cosmological dynamics. In the years that
followed new data have been collected, new methods have been developed, and
new discoveries have been made.
This book describes the presently available observational data on the distribution of galaxies and the application of spatial statistics in cosmology. It provides
a detailed derivation of the basic statistical methods used to study the spatial distribution of galaxies and of the cosmological physics needed to formulate the
statistical models.
We have delineated the basic ideas and practical algorithms and have cited original articles for a more detailed description: there is always more information in
articles than can be collected in a book. We have tried to select articles that present
the subject in the clearest possible way (frequently they are review articles); thus,
our selection is not meant to give a full history of the development of the methods.
Cosmological statistics is a rapidly developing field. We have tried to give the
most up-to-date results (at least up to the year 2001 at the preprint level). After
working through this book, the reader should be able to understand new research
articles and to set up her/his own research projects.
The book is meant to appeal to two different communities of scholars:
graduate students in cosmology and practicing cosmologists
mathematicians interested in methods of spatial statistics and their applications
12
Fractal methods have become very popular in the analysis of galaxy clustering
and Chapter 4 is devoted to this subject. In Chapter 5 we review the history of
the field, introducing the NeymanScott processes and other related geometrical
models. We also give special attention to the Voronoi models and to a physically
motivated Saslaw distribution function. In Chapter 6 we discuss the dynamics of
structure formation, which is important background for formulating contemporary
statistical models of galaxy clustering. The theory of random density and velocity
fields is described in Chapter 7, with emphasis on its cosmological applications.
We focus our attention on both Gaussian and non-Gaussian random fields, with
special attention to the properties and clustering of peaks. We end this chapter by
explaining the mass (intensity) functions predicted by the PressSchechter theory
and the recent halo model of galaxy clustering that is close to the ideas of Neyman
and Scott.
Chapter 8 describes the statistical measures of clustering in Fourier space, in
particular methods of estimation of the power spectrum from observational data.
In the last sections, the properties of lower dimensional fields are discussed, studying what can be inferred from the projected catalogs as well as from one- or
two-dimensional surveys (pencil-beams or slices). In Chapter 9 we explain the
methods of reconstructing the density field in the nearby universe (cosmography),
trying to remove velocity distortions and making use of the statistical knowledge
of velocities and positions of galaxies. We briefly review the application of gravitational lensing for studying the large-scale structure of the universe. In the last
chapter we present the statistical tools that have been developed to highlight morphological features of the galaxy distribution. The topology of the galaxy distribution, used to judge whether we have a cellular or a sponge-like distribution of
galaxies, can be found here together with other morphological descriptors, such
as the Minkowski functionals, minimal spanning trees, and wavelets. Algorithms
for finding clusters, voids, and possible periodicities in the galaxy distribution are
explained in detail. Appendix A gives a short introduction to spherical astronomy and to the coordinate systems used in the catalogs of galaxies, with some
practical formulae to perform coordinate transforms. Appendix B provides a brief
summary of the basic statistical terminology.
13
Acknowledgments
This book has benefited from valuable discussions and, in some cases, from fruitful collaborations with our colleagues, some of whom read drafts of the book and
made useful suggestions for improving it. In particular, we must mention Rien van
de Weygaert, who was one of the originators of the idea for the book and who participated in creating its overall structure, Guillermo Ayala, Adrian Baddeley, Stefano Borgani, Peter Coles, Rosa Domnguez-Tenreiro, Matthew Graham, Bernard
Jones, Martin Kerscher, Sabino Matarrese Belen Lopez-Mart, Chris Miller, Rana
Moyeed, Jesper Mller, Jose Antonio Munoz, Mara Jesus Pons-Bordera, Silvestre Paredes, John Peacock, Jim Peebles, Nurur Rahman, Jose Luis Sanz, JeanLuc Stark, Dietrich Stoyan, Michael Strauss, Istvan Szapudi, Max Tegmark, Luis
Tenorio, and Licia Verde.
This book is profusely illustrated. We created some of the figures; others are
from our earlier papers. Most of the illustrations, however, are from works by
other authors previously published in journals or books. Thanks are due to all
those authors and publishers who generously granted permission to us to make
use of this material. Those authors who have contributed in this way are A. Baleisis, M. Bartelmann, C. Baugh, A. Bijaoui, S. Borgani, A. Canavezes, P. Coles,
M. Colless, L. Christensen, G. Christianson, S. Colombi, R. Croft, L. da Costa,
P. de Bernardis, A. Dekel, V. de Lapparent, J. Einasto, G. Efstathiou, H. El-Ad,
A. Fairall, E. Falco, H. Feldman, K. Fisher, G. Goldhaber, L. Guzzo, A. Hamilton, S. Hatton, W. Hu, G. Kauffmann, M. Kerscher, J.-P. Kneib, O. Lahav, H. Lin,
D. Malin, H. McCracken, S. Maddox, C. Miller, S. Moody, B. Moore, V. Muller,
A. Nusser, J. Peacock, J. Peebles, S. Perlmutter, W. Percival, E. Pierpaoli, T. Piran, D. Pogosyan, V. Sahni, W. Saunders, R. Scaramella, W. Schaap, D. Schlegel,
I. Schmoldt, R. Scoccimarro, T. Souradeep, U. Seljak, E. Slezak, F. Sylos-Labini,
A. Szalay, I. Szapudi, M. Tegmark, M. Turner, H. Valentine, R. van de Weygaert,
L. Van Waerbeke, E. Vishniac, M. Vogeley, S. Webb, D. Weinberg, M. White,
N. Wright, K. Wu, N. Yoshida, and S. Zaroubi.
We have tried to provide a balanced overview of the statistical techniques currently used to quantify galaxy clustering. It has been necessary to deal with a huge
amount of written information. We are grateful to those who created and maintain
ADS and ArXiv, the marvelous archives for electronic communication of research
results.
14
A few parts of the book rely upon, or are extended versions of, some of our earlier published papers, cited in the reference list. We are grateful to the editors and
publishers of the relevant journals and proceedings for permission to use the material: The Astrophysical Journal, the American Astronomical Society, Monthly
Notices of the Royal Astronomical Society, Blackwell Science Ltd., Astronomical
Society of the Pacific, Societ`a Italiana di Fisica, and Springer-Verlag.
We have made every effort to obtain permissions to use copyrighted material.
We apologize for any errors or unintentional omissions, and would be grateful if
they were communicated to us.
This book would not have been possible without the support of many institutions. V. Martnez acknowledges the support of Spanish Direccion General de
Ensenanza Superior Project PB96-0797 and Ministerio de Ciencia y Tecnologa
Project AYA2000-2045. V. Martnez also thanks the Instituto de Astrofsica de
Andalucia and the Observatoire de Gen`eve where parts of the manuscript were
written, and the University of Valencia for providing financial support for these
visits. E. Saar spent several months at the Department of Astronomy and Astrophysics of the University of Valencia, during which time large portions of the
book were written. He is grateful to the Department for its hospitality and for
the invited professor positions funded by the Vicerrectorado de Investigacion de
la Universitat de Val`encia and by the Conselleria de Cultura, Educacion y Ciencia de la Generalitat Valenciana. He also acknowledges the Estonian Academy of
Sciences for a monograph grant and the Estonian Science Foundation for support
in the form of grant 2882.
We thank our editor, Kirsty Stroud, and her assistant, Sharon Taylor, for their
continued interest in the progress of our book, and our production editor,
Christine Andreasen, for her encouragement and, particularly, for her efforts to
translate our Spanish and Estonian English into proper English.
15
CHAPTER 1
1.1 Galaxies
1.1.1 The Milky Way Galaxy
Our Galaxy is a rather flat structure in which we can distinguish three different
parts: the nuclear bulge, in which the galactic nucleus lies, the disc, and the halo
(see Fig. 1.1).
16
2002 by CRC Press LLC
Figure 1.1 Sketch of the Milky Way Galaxy indicating the position of the sun.
The natural unit of length for describing galaxies is the kiloparsec (kpc): 1 kpc
= 1,000 pc ' 3:086 1019 m ' 3; 261:6 light years.
The disc has a diameter of around 30 kpc and is about 0.7 kpc wide. The disc,
with a spiral structure, is populated by young metal-rich stars, dust clouds, and
interstellar gas. The sun lies in the inner part of a spiral arm at 8.5 kpc from the
galactic center. The sun rotates around the center of the Galaxy at a velocity of
220 km s 1 .
The nuclear bulge at the central regions of the Galaxy is 5 kpc wide. The stars
in this region are older than the disc stars. Close to the galactic center lies a strong
radio source (Sagittarius A? ) that could be related to a supermassive black hole.
A spherical halo with a diameter of around 50 kpc surrounds the nuclear bulge
and the disc. The halo is populated with globular clusters, concentrations of thousands of old stars forming very dense and nearly spherical structures. A larger
dark halo surrounds the visible part of the galaxy. Its mass and extent are not yet
completely known.
An Aitoff projection of the Milky Way in galactic coordinates is shown in
Fig. 1.2.
17
2002 by CRC Press LLC
Galaxies
Figure 1.2 A view of our own Galaxy in galactic coordinates (see Appendix A). The center
of the Milky Way lies at the center of the diagram. The interstellar dust prevents a clear
view of the shape of the galaxy. The two spots on the right part of the southern galactic
hemisphere are the Magellanic clouds. (Courtesy of Lund Observatory.)
n = 10
:
Fig. 1.5 shows how we can classify an elliptical galaxy as one type or another,
depending upon the orientation from which the galaxy is observed.
18
2002 by CRC Press LLC
c Gale E. Christianson,
Figure 1.3 Hubbles morphological classification of galaxies. (
reproduced with permission.)
Figure 1.4 On the left we can observe the giant elliptical galaxy M87, in the Virgo cluster.
This is a nearly circular elliptical galaxy classified as an E1 galaxy with a diameter of
about 90 kpc. On the right we can see a wide field view of the Virgo cluster (M87 lies at
c Anglo-Australian Observatory. Photographs
the lower left corner of the photograph). (
by David Malin, reproduced with permission.)
The S0 galaxies or lenticular galaxies are midway between ellipticals and spirals. Like spiral galaxies, they have a nuclear bulge surrounded by a flat disc, but
no spiral arms.
There are two kinds of spiral galaxies, normal and barred. In normal spirals
the spiral arms originate from the nuclear bulge, while in barred spirals the arms
appear at the end of a bar crossing the nucleus itself. Within each group of spirals
several types can be distinguished according to their overall shape. They are referred to as Sa, Sb, and Sc for normal spirals and SBa, SBb, and SBc for barred
19
2002 by CRC Press LLC
Galaxies
Figure 1.5 Elliptical galaxies are classified as a function of their projected image. An
ellipsoidal-shaped galaxy with semiaxis b = c and a > b, seen from A, is classified as
E3, while an observer at B would consider it an E0 galaxy. (Adapted from a figure of B.W.
Carroll and D.A. Ostlie, An Introduction to Modern Astrophysics, Addison-Wesley, 1996.)
spirals, although intermediate types also exist. In both sequences the latter types
have a relative smaller nuclear bulge, thinner spiral arms, and a more open overall shape as shown in Fig. 1.3. Fig. 1.6 shows the images of three normal spiral
galaxies following the sequence Sa, Sb, and Sc. In addition, those galaxies that
cannot fit well into spirals, lenticulars, or ellipticals, having an amorphous shape
without any kind of symmetry, are known as irregulars.
The morphology of galaxies in relation to their spatial distribution allows us to
consider the distribution of galaxies as a marked point process where the mark is,
in this case, a qualitative one. Galaxies of different morphological types present
different degrees of clustering. This property is known as morphological segregation (Giovanelli, Haynes, and Chincarini 1986; Binggeli, Tarenghi, and Sandage
1990). In the astronomical literature elliptical galaxies are known as early-type
galaxies and spirals as late-type galaxies. This is because Hubble thought that
galaxies evolved in his diagram following a sequence from left to right. This is
not the case, but the terminology is still in use. Therefore, it is common to find
statements such as early-type galaxies cluster more strongly than late-type ones.
Galaxies have varying luminosities of a wide range of values, from the luminosity of a dwarf elliptical galaxy of 3 105 L to 1012 L corresponding to
the supergiant ellipticals (L is the luminosity of the sun). The luminosity can
be regarded as a quantitative mark. Luminosity segregation is an observable effect when analyzing the clustering of galaxies with different luminosity. It seems
20
2002 by CRC Press LLC
Figure 1.6 Three spiral galaxies: at the left, M65, an Sa galaxy; at the center, M66, Sb;
c Anglo-Australian Observatory. Photographs by David Maand at the right, M100, Sc. (
lin, reproduced with permission.)
that brighter galaxies are more clustered than fainter ones, although this matter
is still controversial (Alimi, Valls-Gabaud, and Blanchard 1989), and depends on
the luminosity cut and on the analyzed scale. It seems, however, that morphological segregation and luminosity segregation are independent characteristics of the
galaxy distribution (Domnguez-Tenreiro and Martnez 1989).
1.1.3 Brightness and magnitude systems
An important characteristic of a galaxy for a statistician is its brightness this
is frequently used for selecting subsamples from an observational catalog. Astronomers use a historically defined logarithmic scale for describing the brightness of luminous objects if the total observed energy flux from a galaxy is S ,
then the galaxy is assigned an apparent magnitude m by
(1.1)
where the constant defines the zero point of the magnitude scale. Note that the
brighter a galaxy, the smaller its apparent magnitude. The absolute magnitude M
of a luminous object is defined as its apparent magnitude if it were at a distance
of 10 pc. The radiation produced by a galaxy with bolometric (total) luminosity
L emitted isotropically in all directions arrives to an observer at a distance d with
a flux density S . Therefore,
S=
L
;
4d2
(1.2)
and hence the relation between its apparent magnitude and its absolute magnitude
is
m M = 5 log10 (d=1 pc) 5:
(1.3)
21
2002 by CRC Press LLC
Galaxies
The value of m M depends only on the distance to the galaxy and is known as
the distance modulus. Expressing distances in Mpc, we can write
(1.4)
Figure 1.7 In this diagram, we show the annual parallax of a star, P , at a distance, d, from
the sun, where a denotes the semimajor axis of the Earths orbit.
radians, the distance d = a=P . In Fig. 1.7, we see that P is the angle subtended
by 1 AU as seen from the star. This allows us to define the most important scale
unit in astronomy and cosmology, the parsec. One parsec (pc) is the distance at
which a star would present an annual parallax of one arcsecond and therefore 1
pc ' 206,265 AU ' 3.26 light years ' 3.086 1016 m.
The first parallax of a star was measured by F.W. Bessel in 1838. It was for the
star known as 61 Cyg and its annual parallax was 0:300 . The nearest star, after the
sun, is Proxima Centauri, whose parallax is 0:7600 , lying therefore at a distance
of 1.31 pc from the sun. By means of this method, astronomers have measured
distances of about 1,000 stars up to 20 pc from ground-based observatories. Nevertheless, the space mission Hipparcos has already measured the annual parallax
of about 100,000 stars within a distance of 200 pc and with an accuracy of 0:00200 .
Moving clusters
Open clusters are groups of stars formed more or less at the same time and linked
together by their mutual gravitational force. The stars within the cluster move
across the Galaxy with the same speed and pointing in the same direction. These
parallel trajectories, due to an effect of perspective similar to parallel railways
converging at the horizon, seem to converge on, or diverge from, a single point of
the sky. Once we know the position of the convergent point we can determine the
distance to the cluster, first measuring the radial velocity (along the line of sight)
of a star from its Doppler shift vr (in km s 1 ), and then applying the equation
v tan
;
d= r
4:74
23
2002 by CRC Press LLC
Galaxies
Figure 1.8 The real motions of an open cluster are shown as parallel arrows. The angle
between the velocity of a star and its radial component is just , the angle between the star
and the convergent point.
where is the angle between the star and the convergent point and is the proper
motion in the standard units of seconds of arc per year. The factor 4.74 in Eq. 1.1.4
matches the different units used for the angular velocities. With this method astronomers have measured distances up to 100 pc, the most important of which is
the distance to the Hyades cluster in the constellation of Taurus. The distance to
this cluster is 45:75 1:25pc.
Main sequence fitting
In the early twentieth century, the Danish astronomer E. Hertzsprung and the
American astronomer H.N. Russell discovered independently a strict correlation
between the spectral type of a star, which is an indicator of its effective surface
temperature, and its absolute magnitude. This is known as the main sequence of
stars in the HR diagram. Knowing the distance to the Hyades cluster, we deduce
the absolute magnitudes of each star, and we can plot in an HR diagram the magnitude versus the spectral type. This is a calibrated HR diagram. Looking at a
remote cluster of stars, we can now plot on an HR diagram the apparent magnitudes of the observed stars against their spectral types. Since we do not know
the distance to the cluster, we therefore ignore the absolute magnitude. Now we
shift this diagram vertically until its main sequence matches the calibrated one
(see Fig. 1.9). The difference between the absolute and the apparent magnitude of
the two vertical axes provides an indicator of the distance to the remote cluster by
means of (1.3).
24
2002 by CRC Press LLC
10
Figure 1.9 By shifting the HR diagram of apparent magnitudes versus spectral type of
a distant cluster to fit to the calibrated main sequence, we can determine the distance
modulus of the remote cluster.
Periodluminosity relation
Stars whose luminosity changes periodically are known as pulsating variable
stars. One of the best-known types is the Cepheid variable. The name comes from
the star -Cephei. This star changes its brightness following a cycle of variation
of 5.34 days, as shown in Fig. 1.10. In 1912, H. Leavitt found a strict correlation between the period of variability of the Cepheids, P , and their absolute
luminosity, L. This is the so-called periodluminosity relation, which is basically
log L / log P . Fig. 1.10 shows a modern version of this relation for different
types of variable stars. Therefore, we can deduce the absolute luminosity of a
variable star from its period, and hence its distance, by measuring its apparent
magnitude and applying (1.3). Hubble used this method to establish the distance
to the Andromeda galaxy. Recently, the Hubble Space Telescope has made it possible to measure distances to Cepheid variables in other galaxies up to 20 Mpc.
Other distance indicators
Like Cepheids, there are other astronomical objects whose intrinsic brightness
is known. They are generically named standard candles. Apart from Cepheids,
other good standard candles are globular clusters, the most luminous supergiant
stars, HII regions, and type Ia supernovae. For nearby galaxies several independent methods can be applied and astronomers can cross-check their results or take
averages. For more remote galaxies, this is not the case, and in most cases only
the brightest standard candles, type Ia supernovae, are observable.
25
2002 by CRC Press LLC
Galaxies
11
Figure 1.10 On the left, we show the variability of apparent brightness of -Cephei. On the
right, the periodluminosity relation for different kinds of variable stars has been plotted.
12
elliptical galaxies, is
Re / 1:36 Ie 0:85
The hot Big Bang model predicts a universe very dense and hot in the early
stages. In this primordial fireball, matter and radiation rest in thermal equilibrium,
and the universe is opaque. As a result of the expansion, the universe cools down.
When the temperature drops below 4,000 K, photons produced in early phases
from matterantimatter annihilation can freely escape and the universe becomes
transparent. This is the surface of the last scattering that we can observe because
when we look toward the edge of the observable universe, we are looking back in
time, as a consequence of the finite speed of light. The CMB, therefore, represents
our view of the universe when it was about 300,000 years old. This image of the
early universe is very important to understand the development of the large-scale
27
2002 by CRC Press LLC
13
Velocity [km/sec]
30000
20000
10000
0
0
100
200
300
Distance [Mpc]
400
500
Figure 1.11 On the right, we see a modern version of Hubbles law using data from Type
c Edward L. Wright (UCLA),
Ia SNe obtained by Riess, Press, and Kirshner (1996). (
reproduced with permission.) Once the recession velocity of a galaxy has been measured, as for example, for the elliptical galaxy, NGC 4881 (HST image by W. Baum,
AURA/STScI/NASA), we can build a three-dimensional map of the distribution of galaxies, using velocity as indicator of distance. This is shown in the right panel, where this
galaxy, with a radial velocity of 6,691 km s 1 , is placed into the first published slice of the
CfA2 catalog (de Lapparent, Geller, and Huchra 1986).
structure. This fossil radiation was predicted by a group of physicists (R.H. Dicke,
P.J.E. Peebles, P.G. Roll, and D.T. Wilkinson) at Palmer Physical Laboratory in
Princeton, New Jersey, more or less simultaneously with its detection by A.A.
Penzias and R.W. Wilson at Bell Laboratories.
The main features of the CMB are the black-body radiation spectrum and its extreme isotropy. As a consequence of the latter property the universe, at very large
scales, must be smooth. In fact, Ehlers, Geren, and Sachs (1968) showed that the
CMB isotropy combined with the assumption that the Earth is not placed in any
privileged place in the universe the so-called Copernican principle implies
homogeneity. The observations of the distribution of galaxies both in projected
maps and redshift surveys show, however, that the universe is clumpy. Galaxies
and groups of galaxies have a clear tendency to cluster. The theories of structure
formation need to account for both observed opposite trends, namely, the largescale homogeneity and the small-scale clustering of galaxies. The scales probed
by the deepest redshift surveys available today are too small to fill the gap unambiguously, but a gradual tendency to homogeneity has been also detected (Wu,
Lahav, and Rees 1999; Martnez 1999).
The measurements of the tiny deviations of the isotropy in the temperature of
the CMB provided by the Cosmic Background Explorer (COBE) satellite are of
the order of 10 5 measured on scales of about 1000 Mpc, corresponding to the
28
2002 by CRC Press LLC
14
large angular scales at which the satellite was operating. The measurement of the
CMB temperature fluctuations by the COBE satellite in the 1990s provided a fundamental insight into the inhomogeneities of the universe at very early stages.
These anisotropies originated from small perturbations in the gravitational potential. They provide information about the inhomogeneities at the linear evolution
stage, avoiding most of the complications associated with the study of the nonlinear evolution of structures.
Since the discovery of the anisotropies, several statistical analyses of the data
have been proposed. The most widely used ones are the two-point correlation
function and the angular power spectrum. These statistics contain a complete description of the anisotropies if they are Gaussian distributed, and thus are the
most natural choices to test cosmological models leading to Gaussian primordial
fluctuations. However, there are possible physical mechanisms that produce nonGaussian distributed anisotropies. Motivated by this possibility, other statistical
analyses of the anisotropy maps, which are not shaped for Gaussian fields, have
been developed. These include the study of higher-order correlation functions,
morphological characteristics of hot and cold spots, genus and density of spots,
Minkowski functionals, wavelets, and multifractal studies. Some of these techniques are rather similar to the ones this book examines, although in the context
of the galaxy clustering. Of course, they have been adapted to the special character of CMB data, a topic that is beyond the scope of this book. Instead, in the
following sections, we shall concentrate on the clustering of objects, describing
the main galaxy catalogs and the qualitative features observed in them.
15
Figure 1.12 An example of a finger-of-God feature in redshift space around the Coma cluster. (Reproduced, with permission, from Christensen 1986, Compound Redshift Catalogues
and Their Applications to Redshift Distortions of the Two-Point Correlation Function, University of Copenhagen.)
vrec = cz = H0 r + vpec ;
and therefore we have to distinguish between redshift space and real space.
The former is artificially produced by setting each galaxy at the distance r obtained by considering vpec = 0, and is therefore a distorted representation of the
latter. The effect of this radial distortion is clearly illustrated when dense clusters
of galaxies, almost spherical in real space, appear as structures elongated along
the line of sight in redshift space. These structures, known as fingers-of-God,
are well-known features of redshift surveys. A cone around the Coma cluster (see
Section 1.5.1) is highlighted in Fig. 1.12. In the plot, a Cartesian frame has been
displayed, with the observer at the origin. The vertical axis points to the north
30
2002 by CRC Press LLC
16
Real space:
Redshift space:
Squashing effect
Linear regime
Collapsed
Turnaround
Collapsing
Finger-of-G o d
Figure 1.13 A schematic representation of the distortion caused by the peculiar velocities
produced by an overdensity in redshift space. (Reproduced, with permission, from Hamilton
1998, in The Evolving Universe, Kluwer Academic Publishers.)
17
E (B V ) = pD;
where B V is the color index, D is the 100m flux in MJy/sr, as given in the
map, and p = 0:0184 0:0014. The maximum infrared flux in the map gives the
extinction E (B Y ) 0:55, and the extinction toward galactic poles is E (B
V ) 0:015, averaged over a 10 patch. The magnitudedistance relations tell us
32
2002 by CRC Press LLC
18
Dust
270
270
180
180
90
0.33
MJy/sr
Log scale
30.
90
Figure 1.14 Our windows to the universe; the distribution of dust in galactic coordinates
(courtesy of D. Schlegel, D. Finkbeiner, and A. Kriegel). The left panel shows the northern
galactic hemisphere and the right panel shows the southern galactic hemisphere (Lambert equal-area projections). (Reproduced, with permission, from Schlegel, Finkbeiner, and
Davis 1998, Astrophys. J., 500, 525553. AAS.)
nonuniformity of the apparent magnitude-limited samples must be taken into account when performing the statistical analysis of the sample. To analyze this kind
of flux-limited sample we can follow one of two strategies:
1. We can extract volume-limited samples by fixing a value of the depth Dmax
(in h 1 Mpc, h being the Hubble constant in units of 100 Mpc 1 km s 1 ). Then,
from (1.3), we can see that keeping only galaxies brighter than the absolute magnitude limit
Mlim = mlim 25 5 log10 (Dmax ):
(1.5)
the resulting sample is uniformly selected in the sense that the gradient in number density of galaxies with distance because the flux limit is no longer present
(see Figs. 1.15 and 1.16). Dmax should be a luminosity distance (see Chapter 2).
Mlim could be affected by other factors such as the galactic extinction and the K correction (see below). A more general expression for (1.5), incorporating those
effects, must be used, specially when Dmax is large (see Eq. 2.36).
Let us illustrate how volume-limited samples are extracted from magnitudelimited catalogs. For example, if the catalog has an apparent magnitude limit
mlim = 15:5 and we want to take as maximum depth of the volume to be studied Dmax = 101:11 h 1 Mpc, only galaxies with absolute magnitude M
19:70 + 5 log h will remain in the volume-limited sample. With this strategy,
however, we disregard a huge part of the hard-earned information contained in
33
2002 by CRC Press LLC
19
Figure 1.15 This illustration shows how, from a magnitude-limited sample (top panel),
we select a volume-limited sample (bottom panel). The intrinsic brightness of a galaxy is
represented by gray scale (darker gray means brighter). In the magnitude-limited sample
only intrinsically brighter galaxies are seen at large distances. This creates a gradient of
the number density of galaxies with distance. To draw a volume-limited sample, we first
have to find the absolute magnitude limit corresponding to the depth of the sample and
the apparent-limit cutoff (see Eq. 1.5), and then we have to keep in the sample only those
galaxies with M Mlim .
the parent magnitude-limited survey (see Figs. 1.16 and 1.17). To avoid this problem we can follow the second strategy.
2. The second procedure is based on the knowledge of the selection function
'(x). The function '(x) gives an estimate of the probability that a galaxy more
brilliant than a given luminosity cutoff, at a distance x, is included in the sample.
If the sample is complete up to a distance R, '(x) = 1 for x R.
The selection function is derived from the luminosity function (L). The luminosity function is defined by requiring that the mean number of galaxies per unit
volume with luminosity in the range L to L + dL is (L)dL, independent of the
34
2002 by CRC Press LLC
20
Figure 1.16 The left panel shows part of the CfA2 redshift survey. It is a magnitude-limited
sample. All the galaxies in this panel have apparent-magnitude m 15:5 and the catalog
is complete up to this limit. There are 4,933 galaxies. The right panel shows a volumelimited sample with depth 101.11 h 1 Mpc, with only galaxies brighter than 19:70 +
5 log h remaining in the sample. Now there are only 905 galaxies.
location of the volume. The empirical luminosity function is often fitted by the
analytical expression (Schechter 1976; Felten 1977)
L
L
L
exp
d
;
(L)dL =
(1.6)
L
L
L
where L and are the fitting parameters, with related to the number density
(M )dM = A 10
where A =
exp
10
dM ;
(1.7)
is the absolute magnitude for which the catalog is complete. For the Schechter
luminosity function and when the parameter > 1, the selection function can
be written as
( + 1; 100:4(M M (x)) )
;
'(x) =
( + 1; 100:4(M Mmax ) )
where is the incomplete gamma function. Within this strategy, we can assign to
each galaxy a weight w = 1='(x) depending on its distance x from us.
There are different methods to estimate the luminosity function of a galaxy
sample. The most popular method is the maximum likelihood approach, presented
in Yahil et al. (1991). We start with writing the conditional probability density
that a galaxy at a given distance xi has the observed absolute magnitude Mi . It
is given by the luminosity function (Mi ), normalized by the integral over all
35
2002 by CRC Press LLC
21
-10
Absolute magnitude
-12
-14
-16
-18
-20
-22
0
20
40
60
r(z) (h
-1
80
100
Mpc)
500
400
Number
300
200
100
0
0
20
40
60
80
100
Figure 1.17 The top diagram shows the absolute magnitude versus distance for all galaxies in the apparent-magnitude limited survey shown in Fig. 1.16. Only galaxies below the
horizontal line are bright enough to be members of the volume-limited sample. The bottom
panel shows a histogram with the number of galaxies observed as a function of their distance. The huge concentrations around the Virgo and the Coma clusters are clearly shown.
The dashed smooth line indicates the expected number of galaxies we would observe if
they were distributed uniformly in the apparent-limited magnitude sample, calculated by
integration of the luminosity function provided in de Lapparent, Geller, and Huchra (1989).
36
2002 by CRC Press LLC
22
absolute magnitudes it could have at that distance, given the magnitude limit of
the survey:
8
(Mi )
>
< R
;
M (xi )
f (Mi jxi ) =
(
M
)
dM
>
: 0; 1
Mi M (xi );
Mi > M (xi ):
(1.9)
To find the luminosity function we minimize the log-likelihood for all sample
galaxies to have the observed absolute magnitudes
L=
X
i
ln f (Mi jri ):
37
2002 by CRC Press LLC
23
.1
-3
.01
.001
.0001
-5
10
-6
10
-23
-22
-21
-20
-19
absolute magnitude M
-18
-17
-16
Figure 1.18 The luminosity function of the Las Campanas redshift survey. (Reproduced,
with permission, from Lin et al. 1996, Astrophys. J., 464, 6078. AAS.)
24
Figure 1.19 An illustration of the Malmquist bias. All nearby standard candles with a
Gaussian luminosity function are included in the sample, but at large distances only the
intrinsically brighter objects will be included. (Reproduced, with permission, from Webb
1999, Measuring the Universe. The Cosmological Distance Ladder, Springer-Verlag in
association with Praxis Publishing.)
is not a single number but rather a spread of luminosities usually well parameterized by a Gaussian distribution function with mean absolute magnitude M0
and dispersion . Let us illustrate this bias with an example: Let us suppose that
we are looking at standard candles within a magnitude-limited sample of galaxies with apparent magnitude limit m = 13, as shown in Fig. 1.19. When we are
close to the magnitude limit, only the brightest cases of the standard candles are
observed. In other words, galaxies in the distant regions of the sample have to
be systematically brighter in order to be included in the sample. The derivation
of the magnitude of this effect was anticipated by the Swedish astronomer K.G.
Malmquist in 1920. A modern discussion of this derivation can be found in Webb
(1999) for a uniform distribution of the candles in the sample, all of which have
the same Gaussian luminosity function independent of the distance. Under these
conditions the bias is 1:38 2 magnitudes.
39
2002 by CRC Press LLC
25
Figure 1.20 The K-correction as a function of redshift and morphological types. Data
adapted from Shanks (1990) and from King and Ellis (1985). (Reproduced, with permission, from Peebles 1993, Principles of Physical Cosmology, Princeton University Press.)
1.3.6 K-correction
The luminosity of galaxies observed at large redshifts is detected at a longer wavelength than was actually emitted. This shift of the galaxy spectrum toward the red
produces a change in the relation between the absolute and apparent magnitudes
of a galaxy and (1.4) becomes
(1.10)
The K-correction depends on the redshift, the galaxy type, and the waveband in
which the observation has been performed. In Fig. 1.20 we show the K-correction
in the blue B band (Shanks 1990; King and Ellis 1985) for different morphological types, obtained from the analysis of the true spectra of these sources.
1.3.7 Velocity corrections
Redshift catalogs usually list the velocity cz of each galaxy. Several corrections
are applied to this quantity to adapt it to a particular reference frame:
1. Heliocentric frame. The Earth moves around the sun in a nearly circular path
in 1 year at a speed of 30 km/s. Thus, velocities have to be corrected for this
motion to have them referred to a heliocentric frame. This is a rather small
40
2002 by CRC Press LLC
26
correction and generally catalogs already list the heliocentric velocities of the
galaxies.
2. Local group frame. The sun rotates around the galactic center, and the whole
Galaxy has a motion within the Local Group (this is a small gravitationally
bound group of galaxies in which the Milky Way lies; details are given in Section 1.5.1). A standard correction of the heliocentric velocities cz for the solar
motion with respect to the centroid of the Local Group is to add to the heliocentric velocities the quantity 300 sin l cos b, where l and b are the galactic
longitude and latitude, respectively, of each galaxy (see Appendix A for definitions of these celestial coordinates).
3. Cosmic microwave background frame. The frame of reference established by
the thermal background radiation is considered an inertial frame, and therefore
some authors correct the velocities listed in the redshift catalogs for our motion
with respect to the rest frame of the CMB. This motion has a velocity of 371
km s 1 in the direction of l = 264:7 and b = 48:2 (Fixsen et al. 1996).
1.4 Current and future galaxy catalogs
1.4.1 The galaxy distribution in projection
Before any distance indicators are applied to an observed galaxy in the sky, its
accurate angular position is recorded. This information is the basis for the compilation of different catalogs of galaxies known as angular surveys. Typically each
entry of this catalog consists of the angular coordinates of a given object together
with its visual magnitude. Very early in the history of modern cosmology these
catalogs were analyzed using statistical techniques. For example, Hubble (1934)
studied the frequency distribution of galaxies in small angular fields and found
that a lognormal distribution was a good fit for them.
The Lick catalog
The compilation of the Lick survey (Shane and Wirtanen 1967) was an important
challenge for the statistical analysis of projected galaxy maps. The pioneering
work of the Berkeley statisticians Neyman and Scott (1952, 1955) was greatly
motivated by the Lick catalog. The survey consists of about 1 million galaxies
counted over photographic plates. The task was enormous if one considers that
digital scanning and high-speed computer resources were not available at that
time. The authors counted the number of galaxies brighter than m 18:9 in cells
of size 100 100 at > 23 . In the representation of Seldner et al. (1977)
reproduced in Fig. 1.21 we can appreciate for the first time structures on scales of
a few tens of Mpc; in particular, the filamentary structure of the cosmic web can
be glimpsed.
41
2002 by CRC Press LLC
27
Figure 1.21 An equal-area projection of the Lick catalog. (Reproduced, with permission,
from Seldner et al. 1977, Astronom. J., 82, 249256. AAS.)
28
around 2 million galaxies brighter than m = 20:5 have been listed. The galaxies
lie within a solid angle of 1.3 steradians around the south galactic cap.
Fig. 1.23 shows the APM catalog using pixels with brightness scaled to the
number of galaxies in it. A galaxy catalog with the brighter objects was published
by Loveday (1996).
1.4.2 The three-dimensional galaxy distribution
During the past two decades systematic collections of redshifts have been compiled following different strategies. Nowadays, use of multifiber spectrographs
increases the number of available redshifts very rapidly. In this section we describe briefly some of the wide-angle redshift surveys used to date and some of
the projects already under way to map the universe.
Center for Astrophysics (CfA2) redshift survey
The first version of this catalog, the CfA1 (Huchra et al. 1983), sampled only the
Zwicky galaxies brighter than MB = 14:5 in a region of the sky with galactic
latitude greater than 40 or smaller than 30 to avoid the zones of the sky with
high galactic extinction. The extension of the CfA1, the CfA2 catalog (see TDC),
was done by measuring redshifts of Zwicky galaxies brighter than MB = 15:5.
The strategy was to complete contiguous slices 6 wide in declination and 9 hours
43
2002 by CRC Press LLC
29
Figure 1.23 The top panel shows the APM catalog. Brighter areas are more populated. The
dark holes are excluded areas around nearby objects. (Courtesy of S. Maddox, W. Sutherland, G. Efstathiou, J. Loveday, G. Dalton, and the Astrophysics Dept., Oxford University.)
The bottom panel shows the projection of the StromloAPM redshift survey. (Reproduced,
with permission, from Martnez et al. 1998, Mon. Not. R. Astr. Soc., 298, 12121222. Blackwell Science Ltd.)
44
2002 by CRC Press LLC
30
Figure 1.24 A combination of the CfA2 redshift catalog with the SSRS2 survey. The limit in
depth of the slices is z = 0:04 and the equatorial coordinates of the northern slice (top) are
8h < < 17h and 8:5 < < 44:5 , while for the southern slice 20:8h < < 4h and
40 < < 2:5 . 9,325 galaxies are shown. The Great Wall with the Coma cluster in
the northern slice and the Southern Wall with PiscesPerseus supercluster in the southern
slice are the most remarkable features. (Reproduced, with permission, from de Costa et al.
1994, Astrophys. J. Lett., 424, L1L4. AAS.)
in right ascension (de Lapparent et al. 1986; Geller and Huchra 1989; Huchra et
al. 1990). Today it covers a large region of both northern and southern galactic
hemispheres, mapping a solid angle of 2.95 sr: 8h < < 17h , 8:5 < < 44:5
in the north with 6,500 galaxies and 20h < < 4h , 2:5 < < 48 in the
south with 4,283 galaxies (see Fig. 1.24).
Southern Sky Redshift Survey (SSRS2)
In the southern hemisphere ( < 0 ), there is no equivalent to the Zwicky catalog.
In fact, the first version of the Southern Sky Redshift Survey (SSRS) was diameter limited. The task performed by L. da Costa and collaborators was to use the
45
2002 by CRC Press LLC
31
32
33
Figure 1.25 The galaxy distribution for the southern slices of the Las Campanas redshift
survey together with the first slice of the CfA2 catalog at the northern hemisphere. Although
the depth of the Las Campanas is four times (in redshift) the depth of the CfA2 slice, the
size of the structures is the same in both samples, contrary to what is expected for an
unbounded fractal. (Reproduced, with permission, from Martnez 1999, Science, 284, 445
c 1999, American Association for the Advancement of Science.)
446;
square degrees of the sky (Folkes et al. 1999). Fig. 1.26 shows the present status
of the survey after the first 2 years of the observing period. More than 150,000
galaxies already have had redshifts measured (see 2dFGRS).
Sloan Digital Sky Survey (SDSS)
This survey will be the largest galaxy redshift catalog ever compiled (Gunn 1995;
Margon 1999). The observations will be carried out using a 2.5 m telescope in
New Mexico totally dedicated to this task. The telescope carries two double fiber
spectrographs with 320 fibers. The solid angle covered by the survey will be
around sr. Redshifts for more than 1 million galaxies will be measured (see
SDSS).
48
2002 by CRC Press LLC
34
Figure 1.26 A 4 slice containing 63,381 galaxies with measured redshift, drawn from the
2dF galaxy redshift survey. This is the largest view to date of the distribution of galaxies in
the universe. The map shows the very long filamentary structures of galaxies, which could
be cross sections of big walls, clusters, and superclusters, and voids up to 60 h 1 Mpc
in diameter. It substantiates the impression that the size of the cosmic structures already
has been attained. (Reproduced, with permission, from Peacock et al. 2001, Nature, 410,
169173. Macmillan Publishers Ltd.)
35
Figure 1.27 HammerAitoff projection of a compilation of redshifts performed by Christensen (1996) in which some of the features of the nearby distribution of galaxies can be
located, as well as the zone of avoidance (ZoA), the location of the Great Attractor (GA)
close to the galactic equator, and several clusters and superclusters. (Reproduced, with
permission, from Christensen 1986, Compound Redshift Catalogues and Their Applications to Redshift Distortions of the Two-Point Correlation Function, University of Copenhagen.)
about 30 members. Three spiral galaxies dominate the group: the Andromeda
galaxy, M33, and the Milky Way. Nearly all the galaxies within the Local Group
lie within a sphere of radius 1 Mpc. In general, groups of galaxies have fewer than
50 members, mostly spirals, except in very dense and compact groups in which a
good fraction of ellipticals can also be found.
Clusters of galaxies are larger systems containing from 50 up to thousands of
members within a sphere of radius up to 4 Mpc. They contain galaxies of all
types. There are clusters with a well-defined spherical shape, but others show a
more irregular and open structure. It is also possible to detect substructure within a
large cluster. One example of an irregular cluster is the Virgo cluster (see Fig. 1.4).
It has an average radial velocity of 1,100 km s 1 , and therefore the center of the
cluster lies at a distance of 11 h 1 Mpc. From Fig. 1.27 we can see that the cluster
covers a vast region of the sky (approximately 10 10 ). Fig. 1.28 shows where
50
2002 by CRC Press LLC
36
Figure 1.28 A map showing the main large-scale features in our local universe. (Reproduced, with permission, from Fairall 1998, Large-Scale Structures in the Universe, John
Wiley & Sons in association with Praxis Publishing.)
this cluster lies in relation to our position. In fact, our Galaxy forms part of one
of the branches growing outward from the cluster. The morphological segregation
between ellipticals and spirals (see Section 1.4.3) is clearly evident in the spatial
distribution of galaxies within the Virgo cluster. The central region of the cluster
is populated mainly by elliptical galaxies. The biggest one, M87, is a giant E1
elliptical (see Fig. 1.4). Spirals are more or less distributed as a halo surrounding
the cluster core.
51
2002 by CRC Press LLC
37
The more famous regularly shaped rich cluster of galaxies is the Coma cluster,
lying 5.4 times farther away from the Virgo cluster (see Fig. 1.28). Of the galaxies
in the Coma cluster 85% are ellipticals and S0s. The diameter of the cluster is
roughly 6 h 1 Mpc although, as a result of peculiar motions, it stretches into an
elongated structure, resembling a stick-man in redshift space (see Fig. 1.12).
Closer to the Earth and very similar to the Coma cluster, the cluster Abell 3627
was discovered later because, with a rather low galactic latitude, it is partially
obscured by the dust of our own Galaxy. Another very big cluster is the Perseus
cluster, which lies in the south galactic hemisphere. It is also partially obscured
by the Milky Way.
1.5.2 Catalogs of clusters of galaxies
Clusters of galaxies are the largest and more massive virialized objects in the
universe. As well-defined cosmological entities, clusters are used to probe the
large-scale distribution of matter in the universe and their statistical properties
complement and are related to those of the galaxy distribution. Similar to the
galaxy catalogs described earlier, cluster catalogs also have been compiled, although some uncertainties regarding their statistical analysis have arisen from the
loose definition of what a cluster is or from subjective visual identification of a
cluster in projected maps (Sutherland and Efstathiou 1991). Projection effects can
easily contaminate the detection of a cluster (van Haarlem 1996). For example, a
filament along a line-of-sight could easily appear to be a rich cluster on a projected
map.
Lately, automatic detection of clusters has been applied using more objective
criteria and scanning machines. There is still a more powerful and less ambiguous
way to detect these massive concentrations of galaxies. The hot intra-cluster gas
produces X-ray emission by bremsstrahlung radiation. The X-ray luminosity is
correlated with the cluster mass, and therefore can be used to infer the cluster
richness, i.e., the number of galaxies in the cluster (see Reiprich and Borhinger
1999). In this way, galaxy clusters can be objectively detected by present-day
X-ray satellites. In particular, the ROSAT satellite has provided more than 1,000
positions of clusters within a huge volume of the universe (Borgani and Guzzo
2001).
In this section we describe some of the cluster catalogs used as tracers of the
large-scale structures of the universe. Statistical study of cluster catalogs is performed using tools similar to those used in the study of galaxy catalogs and explained elsewhere in this book.
The Abell cluster catalog
By visual inspection of the Palomar Observatory Sky Survey (POSS), Abell (1958)
identified 2,712 rich clusters of galaxies. This catalog was the first well-defined
52
2002 by CRC Press LLC
38
and statistically useful sample of clusters of galaxies. The criteria used to select
a cluster from the plates were the following. A cluster must contain at least 50
members with apparent magnitude in the range [m3 ; m3 + 2], where m3 is the
third brightest member of the cluster. All cluster members must lie within a radius of 1.5 h 1 Mpc (Abell radius) from the center of the cluster. The survey
contains clusters lying north of declination = 27 . Abell assigned a richness
class to each cluster depending on the number of members according to the previous criteria. Six richness classes were established, from R = 0 to R = 5. The
number of cluster members of each class varies according to the following table.
Richness class
0
1
2
3
4
5
3049
5079
80129
130199
200299
300 or more
39
40
Figure 1.29 The REFLEX cluster catalog. The zone of avoidance around the galactic plane
( 20 < b < 20 ) is not mapped. (Reproduced, with permission, from Borgani and Guzzo
2001, Nature, 409, 3945. Macmillan Publishers Ltd.)
(1981) and known as the Bootes void, has an overall size of 75 h 1 Mpc. This size
is larger than the size of all other voids observed in the slice redshift surveys (like
the CfA2 or Las Campanas). The maximum void size in these surveys is about
60 h 1 Mpc in diameter. In fact, the Bootes void was found to present some kind
of rarefied structures in its interior. Voids in general are roughly spherical with
some inlying galaxies pointing to their interior, although in many cases the core
appears to be really empty. It is worth mentioning that the sizes of the voids do
not increase with the depth of the redshift survey, as we can see in the deepest
slice surveys constructed thus far, such as the Las Campanas redshift survey or
the 2dF (see Figs. 1.25 and 1.26).
1.5.5 The texture of the galaxy distribution
We conclude this chapter by summarizing the overall view that the large-scale
structures in the universe depict. Different expressions are often used to describe
these large patterns: cellular structure, sponge-like topology,z foam-like bubbles
z A mathematical motivation for the use of sponge-like topology is given in Chapter 10.
55
2002 by CRC Press LLC
41
56
2002 by CRC Press LLC
CHAPTER 2
44
system with a universal cosmological time, so that we can decompose the metric
into its temporal and spatial parts, and homogeneity and isotropy together demand that the space-like sections must be spaces of constant curvature. (In fact,
requiring isotropy for every observer implies homogeneity.) We know from observations that the universe is expanding, so we have to allow this curvature to
change in time.
While the cosmological principle is certainly philosophical, its consequences
and the principle itself can be tested, if we manage to observe a fairly large region
of the universe. We shall describe the present status of these direct tests below.
The geometry of the four-dimensional space-time is described by its metric,
ds2
3 X
3
X
gij
(x) dxidxj ;
i=0 j =0
where ds is the distance between two nearby points (events) with coordinates
x; x dx and gij x is the metric tensor. The cosmological principle leads us to
the RobertsonWalker metric
()
= c2 dt2
()
R2 t
d! 2
()
()
()
()
The coordinates !; ; are called comoving coordinates, and the scale factor R t
describes the general expansion (or contraction) of the space. Thus, distances
58
2002 by CRC Press LLC
45
()
= R2(t)d!2 ;
d
= d = 0:
The observations we make lie along our past light cone, given, obviously, by
d!
dt
c
:
Rt
()
(2.3)
Strictly speaking, all distances along a light cone are zero, by the definition
of the metric in Riemann geometry. The distances that can be used are the time
intervals (see Fig. 2.1), physical distances (lengths of the arcs from the time axis),
and distances measured by dimensionless coordinates ! .
It would be meaningless to compare physical distances of galaxies because
these depend on the time coordinate. The radial coordinate ! of an object at rest,
on the contrary, does not change with time and can be used to find distances
between objects. Such distances are called comoving distances. A comoving distance can be expressed either in dimensionless units of ! , or in the units of length
(R0 ! ). These distances can be thought of as distances between galaxies at the
59
2002 by CRC Press LLC
46
t
O
2
t1
* 1
t2
* 2
Figure 2.1 Observations along a light cone (thick line). The location of the observer is
marked by O and locations of two luminous objects (galaxies) are indicated by asterisks.
Each galaxy and the observer define their instantaneous three-space (arcs in the figure).
The time and dimensionless radial coordinate of the observer are T and , and for the two
galaxies t1 ; !1 and t2 ; !2 , respectively.
present moment (along the space-like section defined by the age of the universe
T ), although we have observed them only in the past and cannot see them at their
present location.
So, the first step in determining a distance between two galaxies is to find their
comoving coordinates ! (comoving distances from the observer who has chosen
). This depends on the assumed cosmological model and
his coordinates as !
we shall describe it a little later. The second step is to find the distance between
two galaxies in curved three-dimensional space.
; g (the same used for the
Let us denote their coordinates as f!i ; i ; i ; i
line element (2.1)). The coordinates and are the usual spherical coordinates in
the sky ( is the zenith distance and is the longitude). From a spherical triangle
formed by the pole of the coordinates and by the points fi ; i g we can find
the angular distance between the points on the sky by the cosine theorem of
spherical trigonometry:
=0
=12
2 :
(2.4)
Note that here i are zenith distances, measured from the pole of the spherical
coordinate system. Most astronomical coordinate systems use latitudes, measured
from the equator; if this is the case with your catalog, you have to modify the
above formula, replacing ! .
60
2002 by CRC Press LLC
47
12
1
Figure 2.2 Determination of the distance between two events. The spherical triangle is
formed by our location (0) and by the two objects (1,2).
Let us now move the pole of the celestial coordinates to the location of one of
our objects. In the new coordinates
d
= 0;
d
= d;
= R02(d!2 + Sk2(!)d2 ):
This formula tells us that in order to find the distance l between two objects we
have to solve a triangle on an appropriate two-dimensional surface, either a sphere
), a plane (k
), or a pseudo-sphere (k
). The angular distance is
(k
similar to the longitude and the dimensionless radius ! to the zenith distance
of a usual spherical coordinate system. Fig. 2.2 shows determination of distance
on a sphere.
) we have the cosine law:
For a flat three-space (k
=1
=0
= 1
=0
2 = ! 2 + ! 2 2! ! cos :
!12
(2.5)
1 2
1
2
For a closed hypersphere (k = 1) spherical trigonometry tells us that
cos !12 = cos !1 + cos !2 + sin !1 sin !2 cos :
(2.6)
For the negative curvature, open universe (k = 1) there exists an analogous
(2.7)
These formulae give the dimensionless distances; in order to get the distance in
units of length, we have to multiply the dimensionless distance by R0 :
r12
= R0 !12 :
61
(2.8)
48
We shall see below that for the Friedmann cosmological models in the case of a
) R0 and ! cannot be defined separately, but their product
flat three-space (k
r
R0 ! can. Multiplying both sides of the formula (2.5) by R02 transforms it
into an equivalent formula for r.
=0
Volumes
According to Riemann geometry, the spatial volume element dV can be written
as
p
dV
g d3 x;
det
where g is the spatial part of the metric tensor. For our metric it gives
dV
(2.9)
The corresponding comoving volume (the volume delimited by the same markers
at the present epoch) is given by the same formula, if we substitute R0 for R t .
The volumes of the observed galaxy samples are usually spherical sections.
The comoving volume of a sphere with a radius ! is
()
V !
= 1;
k = 0;
k = 1:
(2.10)
( )
R_ (t)
dr
=
R_ (t)! =
r = H (t)r
v=
dt
R(t)
(2.11)
is proportional to their distance. This relation is called Hubbles law galaxies move away from us at velocities that are proportional to their distance from
us. This law was established observationally by E. Hubble in 1929. For nearby
galaxies and for the present epoch
v
The ratio
= H0 r:
_
( ) = RR((tt)) = aa_ ((tt))
H t
is Hubbles function. Its value at the present epoch H0 is called Hubbles constant
and must be found from observations.
We do not observe directly the recession velocity of a galaxy, but the changes in
its spectra. Because of the Doppler effect, the light emitted from receding galaxies
62
2002 by CRC Press LLC
49
= 1 +ev=c ;
0
where v is the velocity of the emitter and c is the velocity of light (strictly speaking, this formula is valid only for small velocities, but we are using it for such a
case). For small velocities (and small changes of frequency) we can write
d
v
c
_( )
R_ (t)
=
dt =
()
R(t)
R t r
R t c
dR
:
R
1=R(t);
= 2 R(t):
(2.12)
The Doppler interpretation of the change of the wavelength of radiation is useful for nearby objects and small velocities. For far-away galaxies the notion of
their speed in respect to us has no clear meaning, because their light was emitted
so long in the past. The formulae (2.12) remain correct in this case, too. In order
to see that, we can imagine one standing electromagnetic wave (radiation of a single frequency) in an expanding universe. During expansion the number of nodes
of this wave cannot change; thus, the wavelength of the wave has to grow at the
same rate as the scale factor R t .
The observed shift of the frequency (or wavelength) of the spectra is described
by a parameter z called redshift:
()
= 0
e
and this is the measure of distance we can find in observational catalogs. Here
e is the wavelength of a spectral line at the moment of emission (we know it on
the basis of studies of spectra of nearby galaxies), and 0 is the wavelength we
observe that line at here.
This definition, together with the formula (2.12) above, gives us
1 + z = RR(0t) = a 1(t):
(2.13)
Thus, measuring the redshift of a luminous object, we can determine the scale
factor R t at the moment when the light was emitted. If we know the R t dependence, we can find the cosmological time t, and from the equation of the light
cone (2.3), the spatial distance to the object ! .
()
()
63
2002 by CRC Press LLC
50
gij = 8G
Tij ;
c4
where the Einstein tensor Gij is formed from the metric tensor gij , its first and
second derivatives, and is the cosmological constant. The energy-impulse tensor Tij on the right-hand side of the equation describes the physical content of the
universe, the coefficient of proportionality given by a combination of the gravitational constant G and the velocity of light c.
For the RobertsonWalker line element (2.1) the above equations reduce to the
Friedmann equations:
scale factor, is the total energy density, p is the pressure and k is the curvature
constant (see 2.2). These equations together with the RobertsonWalker metrics
(2.1) define the standard cosmological models. The three equations (2.142.16)
are not independent; the first equation (2.14) can be derived, using the other two
equations.
The main contributors to the total energy density are baryonic matter, radiation, and nonbaryonic (dark) matter. These all have different equations of state
and they dominate at different epochs of the evolution of the universe. For epochs
when we can count objects and determine their statistics, the energy density of
radiation is negligible compared with that of baryonic matter and dark matter, and
for these constituents the pressure term p=c2 in the Friedmann equations is much
and to
smaller than the density . Thus the usual approximation is to set p
consider the universe to be filled with pressureless dust. In this case we get from
the equation (2.16):
/ R 3:
(2.17)
=0
M = 8G
3H 2 ;
0
(2.18)
where 0 is the present average matter density, and the normalized cosmological
64
2002 by CRC Press LLC
51
constant as
= 3Hc 2 :
0
As the values of the parameters
M ;
determine the evolution of the cosmological model, they are called the cosmological parameters. The combination
(
M = 1;
= 0) defines the Einsteinde Sitter model and (
M =
= 0)
describes the Milne model, an empty expanding universe.
Another important parameter is the Hubble constant H0 , which determines the
overall spatial and temporal scales of the universe.
Cosmological models can be also parameterized by the deceleration parameter
q , which completely determines models without the cosmological constant:
q
:
= RR
R_ 2
=
2M
:
()
( ) = H0 E (z):
H z
()
(2.19)
The best designation for that function would be h z , but h is already used in
h
the same context to designate the dimensionless Hubble constant (H0
km/sec Mpc 1 ). Using (2.17) and (2.13), we can rewrite the Friedmann equation (2.15) as
100
( ) =
M (1 + z)3 +
Hkc2 R2 (1 + z)2 :
0 0
At the redshift z = 0 the function E (0) = 1, giving
kc2
=
M +
1;
H 2 R2
E2 z
0 0
(2.20)
( ) =
M (1 + z)3 + (1
M
)(1 + z)2 +
:
E2 z
(2.21)
This function allows us to describe the evolution of the cosmological model and
to express all the observational relations in terms of an observed quantity, the
redshift z . Some examples of that are given below.
We see also from (2.20) that the type of the curvature k is given by the sign of
the difference M
. This combination of parameters is sometimes called
the curvature parameter K and is defined as
+
1
K = 1
M
:
Thus for negative curvature, k = 1,
K > 0, and for k = 1
K < 0.
65
52
At any epoch z the cosmological model can be described by the current (instantaneous) cosmological parameters M z ; z ; H z . The expression for
H z is given by the formula (2.19) and the formulae for the other parameters
follow from the definition of M ; , above:
()
() ()
()
M (z) =
ME(12 (+z)z) ;
(z) = E
2 (z) :
(2.22)
(2.23)
=0
( ) =
(1
+ (1) ++ 1 )
and we see that initially (z 1) this model behaves as the Einsteinde Sitter
model,
M (z ) 1. This behavior lasts until the redshift zc given by zc + 1
(1
M )=
M , when the model starts expanding as a true open model. For a
reasonable
M 0:3 this happens rather late, at zc 1:3.
For a true k = 0 universe with the same density parameter
M = 0:2,
=
0:8, the fast expansion starts much later, at zc 0:6. For the Einsteinde Sitter
model the fast expansion stage never arrives.
2.2.4 Cosmological time
The expression for cosmological time can be derived from the definition of the
R0 =
z :
function g z (2.19) and the fact that R z
()
( ) = (1 + )
1 dR = 1 dz :
H = H0 E (z ) =
R dt
1 + z dt
(2.24)
When this equation is integrated, the age of the universe at the epoch described
by a redshift z is given as
1
t(z ) =
H
0 z
dz 0
:
z0 E z0
(1 + ) ( )
(2.25)
The time elapsed from the epoch z until now (the look-back time) is given by the
same integral with the limits and z , and we get the present age of the universe,
if we integrate from to 1. This age is finite for all possible models.
For general combinations of M and the integral has to be computed nu, there exist analytic expressions for it, different for different
merically. If
k , but these are very cumbersome and numerical integration is recommended in
that case, too.
=0
66
2002 by CRC Press LLC
53
Simple analytic results exist for two models that can be used for quick esti;
) we get
mates. For the Einsteinde Sitter model ( M
= 2 (3 )
=1
=0
1
2
t(z ) =
3H0 (1 + z)3=2
with T
= H0 for the age of the universe. If we substitute here H0 with the
z 3 ,
present mean density of the universe 0 from (2.18) and note that 0
we get the density behavior in the Einsteinde Sitter model:
(1+ ) =
1 :
= 6Gt
2
(2.26)
This model is not only the simplest one, but it is also an excellent approximation
for all other models for large redshifts. This is easy to understand if we look at the
high-z asymptotics of the functions M z and z :
()
()
M (z) 1 + (
M +
1)z 1 ; z 1
(z)
M z 3; z 1:
For the empty Milne model
M =
= 0
1 1
t(z ) =
H0 1 + z
with T = 1=H0 for the age of the universe. This model can be thought of as
representing a low-density universe (the descriptive functions for
= 0:1, for
example, do not differ much from those for the empty model).
2.2.5 The light cone equation
In order to find the comoving radial distance of an observed event with the redshift
z , we need to solve the light cone equation (2.3):
d!
dt
c
:
Rt
()
Using the expression for dz=dt from (2.24), we can integrate this equation to find
Z z
c
dz 0
!
(2.27)
:
R0 H0 0 E z 0
( )
=
+
( )
67
2002 by CRC Press LLC
, by formula
(2.28)
54
4.5
4
3.5
E
2.5
2
1.5
C
1
0.5
E-S
0
0.1
0.2
0.3
0.4
0.5
z
0.6
0.7
0.8
0.9
Figure 2.3 The volume-redshift dependence for the Einsteinde Sitter model (E-S) and the
concordance model (C). We also show the Euclidean behavior (E) for comparison. The
volumes are given in the units of cubed Hubble length c=H0 3 .
=0 =
=0
= H2c 1 p11+ z ;
M = 1;
= 0;
0
c
= H ln(1 + z);
M = 0;
= 0:
To illustrate this, we show in Fig. 2.3 the growth of the comoving volume of a
sphere (formula 2.10) of a radius of redshift z in the Einsteinde Sitter cosmology
: ;
: . The latter model
and in another (concordance) model with M
is based on a comparison of the recent cosmic microwave background (CMB) data
with data on the large-scale structure of the universe (Tegmark, Zaldarriaga, and
Hamilton 2001). Although the three-space is flat in both those models, the spacetime is curved, the ! z relation is nonlinear, and the volume does not grow as in
the Euclidean case (V z 3 ).
An important fact that can be read from formula (2.28) and the expression for
E z (2.21) is that the integral is finite for infinite z there exists a maximum
comoving distance rmax r z ! 1 (the only exception is the empty model).
This is called the (particle) horizon, and it is caused by the finite lifetime of the
universe; we can observe only those sources whose light has had enough time to
=03
=07
()
()
= (
68
2002 by CRC Press LLC
55
reach us. Thus, we can, in principle, observe only a finite volume patch of the
universe, even if the universe itself is infinite. This is a specific feature of the
standard model we have described here; there are models for the early evolution
of the universe that do not have particle horizons.
2.2.6 Observational distances
There are, by tradition, many distance definitions in observational cosmology.
These are not really distances, but rather recipes for calculating different physical
aspects of light sources.
Let us consider an extended object (galaxy) at an epoch z . Using the expression
, we can find the physical size of the galaxy
for metrics (2.1) and setting d!
l, knowing its angular diameter and comoving distance ! :
=0
l = R(t)Sk (!);
l = D ;
where D is the distance to the galaxy. The comparison of these two formulae leads
to the definition of the angular diameter distance:
= 1R+0 z Sk (!(z)):
Da
So, if we insist on using Euclidean formulae for this procedure, an angular diameter distance is the function of z we have to use to calculate the angular diameter
of an object. This is the recipe for other distance definitions, too.
Interestingly, the function Sk ! z has the same rather simple form for all M
in the case
, although the function ! z itself has to be calculated by three
different formulae. If we define another distance Dm z by
( ( ))
=0
()
( ) = R0 Sk (!(z));
()
Dm z
(2.29)
p
2
c
M z + (
M 2)( 1 +
M z 1)
Dm (z ) =
H0
2M (1 + z)
(2.30)
This formula is known as the Mattig formula, and it sometimes is written in terms
of the deceleration parameter q
M = . For small M it is better to use another
form of it (Peacock 1999):
=
2
( )=
Dm z
c
H0
(1 + z + p1 + p
M z)
(1 + )(1 +
M z=2 + 1 +
M z) :
z
z
( ) = Dm(z)=(1 + z):
Da z
69
2002 by CRC Press LLC
56
1
0.8
distance
Dl
M=0.3, =0.7
0.6
Dm
0.4
Da
0.2
0
0
0.2
0.4
0.6
0.8
=03
=07
= 05
L
= 4D
2:
(2.31)
( ) = (1 + z)Dm (z):
(2.32)
Because the formula for Dm (z ) is so simple, it has been used extensively as
a comoving distance in cosmological statistics of deep surveys. For smaller z
the difference between Dm (z ) and the geometrical comoving distance r(z ) =
R0 ! (z ) is not great. Note that in the case k = 0 (and
= 0) both comoving
distances coincide.
As an example, we show in Fig. 2.4 the distances r(z ); Da (z ), and DL (z )
for the concordance model with parameters
M = 0:3;
= 0:7. The Mattig
distance is not defined for these models, but the usual practice is to use for all
models the Mattig distance formula for q = 0:5. This curve is labeled Dm in the
figure.
The deepest present galaxy surveys reach z 0:3; at these distances the margin
for error if we use the wrong distance definition is about 10%. New surveys will
reach farther, and quasars have been observed up to redshifts z 5, where the
DL z
70
2002 by CRC Press LLC
57
difference between the Mattig distance and the true comoving distance is about
30%. The effects we are trying to measure are much smaller; therefore, we must
be careful when using distance-based statistics in cosmology.
We explained earlier in this chapter that the distance we need in spatial statistics
is the comoving distance. It is the length of the geodesic of a space-like section,
the minimal space-like distance possible, and it is the only counterpart for the
Euclidean distance used in traditional statistics.
The method for calculating distances depends on the cosmological model that
), as all popular models
we assume. If this model has flat spatial sections (k
do, the calculation is simple. As the three-space is flat, we can continue using
Cartesian coordinates. The only change is the mapping from redshift to comoving
distance from the observer, given by Eq. 2.28. The distances between galaxies can
then be found by the usual Euclidean formulae.
If the chosen model has curved spatial sections, the calculations are more complex. Although the spatial sections are homogeneous, coordinate systems that
cover the whole spatial section are singular (as in the case of spherical coordinates), and the usual Euclidean coordinates cannot be used. First, we have to find
the dimensionless comoving distances from the observer (2.27) for all galaxies. If
we have to calculate distances between galaxy pairs, we have to find the line-ofsight angles for every pair by (2.4) and then use these angles to find the dimensionless comoving pair distances using (2.6) in the case of positive curvature or
using (2.7) in the case of negative curvature. Finally, we have to use Eq. 2.8 to
convert to comoving distances.
There is also the possibility of continuing to use Euclidean coordinates for
galaxies, but in an artificial four-dimensional Euclidean (or Minkowski) space
that embeds the spherical (or pseudo-spherical) three-dimensional spatial sections
(Roukema 2001). The distances are then calculated as arc-lengths in four dimensions. The number of calculations is the same as for the method described above,
but the concept could prove useful for future applications.
In most applications of spatial statistics in cosmology one of the goals is to
estimate the cosmological parameters M and . If we vary the values of these
parameters, we should recalculate the galaxy positions and mutual distances for
the initial catalog. Because this is an enormous amount of work, it is usually
avoided. This approach can be justified for comparatively shallow catalogs, but
the margin for error must be estimated.
=0
58
Let us first give some numbers to get the feel of the characteristic scales.
The natural unit of distance in cosmology is the megaparsec (Mpc), Mpc
6 pc
:
24 cm. Average distances between galaxies and sizes of
galaxy clusters are of the order of a few Mpc. As one parsec is 3.26 light-years, a
step of one Mpc in space along the light cone takes us another 3.26 million years
into the past.
We have seen above that the age and the radius of the curvature of the universe are determined by Hubbles constant. The values obtained for it vary rather
widely, but the recent measurements seem to converge in the interval
6070 km sec 1 Mpc 1 . In theoretical papers H0 is usually expressed as H0
h km sec 1 Mpc 1 , giving the explicit dependence of the results on h. For
km sec 1 Mpc 1 (h
: ).
further estimates we shall use the value H0
=H0
The age of the universe (the Hubble age) will then be about T
9 years, and its typical size (the Hubble radius) will be about c=H0 4,600
Mpc. The relative amplitude of the curvature effects is of the order of the redshift
z , or, if expressed in recession velocities, of the order of the ratio v=c (c 5
km/sec is the velocity of light). Another typical length, the curvature radius R0 ,
given by
c
p
R0
;
H0 j M
j
is difficult to estimate if we do not know the cosmological parameters M and ,
and it describes only the curvature effects inside the space-like sections; space; R0 1.
time is well curved even in the case k
Because all distances and the look-back time tl are linear functions of z for
small z , the constants found above can be used to estimate them:
c
d
z 4,600 z Mpc;
z ;
H0
= 3 086 10
10
100
= 65
= 0 65
=1 =
=
3 10
15 10
=0
tl
H1 z 15 109 z yr;
0
1
z 1:
The present mean density of the universe is easy to calculate, using the parameterization formula (2.18) from above:
= 1:9 10 29
M h2 g=cm3 = 2:8 1011
M h2 M =Mpc3;
(2.33)
59
light intensity. Thus, many galaxies, especially those farther away from us, remain
unobserved.
All matter is not in galaxies. Most of it (9599%) is dark and has been detected
only by its dynamical effects. Its physical nature is not yet clear; because it determines largely the evolution of the structure in the universe, the hypothesis about
the dark matter is frequently considered an extra cosmological parameter.
There are also brighter objects that can be observed for larger distances, quasars
and galaxy clusters. A (rich) galaxy cluster contains from a few hundred to a
thousand galaxies and its characteristic size is a few Mpc. The spatial density of
5 Mpc 3 .
galaxy clusters is about
10
L
= 4r
2;
where L is the luminosity of the object. In an infinite Euclidean universe, populated with luminous objects of an average luminosity L with the number density
n, the total observed flux density Stot would be given by the integral
Stot
nL 2
r dr;
r2
which evidently diverges. In fact, we would not see an infinitely bright sky, as
foreground stars would screen those farther away, but there would be a star shining
at any point in the sky. This result is known as Olbers paradox.
Contemporary cosmological models do not predict an infinite luminosity of the
sky for several reasons. First, the existence of the particle horizon means that the
volume we observe is finite. Second, it is reasonable to suppose that in an evolving
universe stars and galaxies form and die, and the spatial density of light sources
will not be constant on the light cone. Third, the expansion of the universe causes
the photons to lose their energy:
E
/ / 1=R(t);
()
where E is the energy of a photon, its frequency, and R t the scale factor.
These factors all work to diminish the brightness of the sky, but they cannot
reduce it to zero. The photons emitted by atoms during the evolution of the universe form the background light, and observations of its intensity and spectra (or
corresponding observational limits) can be used to constrain theories of formation
of structure in the universe.
73
2002 by CRC Press LLC
60
Apart from radiation from discrete objects, there is a component of the background that once really made the whole sky shine uniformly. Because this happened at redshifts before z 1,300, this light by now has been redshifted into
the microwave region of the spectrum ( mm), and it is called the cosmic
microwave background (CMB) radiation.
It can be shown that in an expanding universe the temperatures of matter Tm
and radiation T
depend on the value of the scale factor as
order to determine how hot we have to know the present temperatures. The present
mean temperature of matter is difficult to determine because it has been affected
by many factors other than the cosmological expansion. As the energy density of
z 4 and that of matter /
z 3 , it was the energy
radiation
/ T 4 /
density of radiation that determined the overall evolution of the universe in early
:
K is very small, at
times. Although the present CMB temperature T
the redshifts z > 1,300 the temperature and the energy density of radiation were
high enough to ionize all matter and to keep the matter and radiation temperatures
equal.
As the universe expanded, temperatures dropped and at z 1,300 recombination started, neutralizing the ionized matter and releasing the radiation. Because
the only change that the radiation has undergone is the change of the wavelength,
it has safeguarded information about the properties of the universe before recombination (recall that matter and radiation were then in close interaction).
(1 + )
(1 + )
= 2 74
61
Figure 2.5 Positions of the 15,431 PSCz survey galaxies in the galactic coordinates. The
at the right to l
at the left; the galactic north
galactic longitude grows from l
pole is at the top. The black masked regions either have not been observed or were excluded
from the survey due to a too high infrared sky background. The gray masked regions have
a high galactic extinction. (Reproduced, with permission, from Saunders et al. 2000, Mon.
Not. R. Astr. Soc., 317, 5564. Blackwell Science Ltd.)
=0
= 360
avoidance. Because dust is more transparent to infrared radiation, infrared surveys can achieve a better coverage of the sky.
For example, we show in Fig. 2.5 the sky distribution of 15,431 galaxies with
measured redshifts from the PSCz survey (Saunders et al. 2000), based on the
IRAS satellite survey of infrared objects. This is as complete sky coverage of
galaxy positions as possible for the present time, and even here you can see the
traces of the galactic plane and unobserved tracks (it covers 84% of the sky).
Although foreground structures can be seen in places, the overall distribution of
galaxies is fairly isotropic, showing no preferred directions.
Better results can be obtained using radio sources that are usually more distant and are not affected by the galactic absorption. But the best limits on possible anisotropy can be obtained from the measurements of the background radiation in different regions of spectra. Thus, as described in more detail by Peebles
(1993), the measurements of the X-ray brightness of the sky limit a possible large3 . Even better limits come from the measurement of the
scale anisotropy by
cosmic microwave radiation background, where the temperature fluctuations at
5 (Bennett et al.
scales larger than have been found to be about T =T
1996). To be exact, there is one large-scale feature in the temperature distribution, a dipole moment, but that is due to our movement relative to the microwave
background (the Doppler effect).
Thus, we can consider the assumption of the isotropy of the matter distribution
to be well confirmed by observations.
10
75
2002 by CRC Press LLC
10
62
()
()
(2.34)
The only difference is that we have to replace the distance by the cosmological
luminosity distance DL (2.32). Of course, the larger the volume is, the brighter
and fewer galaxies we have to use.
As an example, Fig. 2.6 shows the number-redshift relation for the very deep
(up to z : ) ESO Slice Project redshift catalog, from Scaramella et al. (1998).
Although there are 3342 galaxies in the catalog, the numbers for the volumelimited samples are much smaller, ranging from 815 to 43 (see the legends). The
number counts are approximated by power laws, N Rn , and the exponents
n are shown in the figure. We see here all the problems of number counts. For
smaller distances, up to z : , the local large-scale structure is distorting the
picture, but at larger distances cosmological effects come into play.
In order to transform from redshifts to distances, we should know the parameters of the cosmological model, which are not well fixed at the moment. Strictly
speaking, at these redshifts one should use the volume-redshift relation, and not
look for a power law.
The second problem is the K-correction, described in Chapter 1. As we see
in Fig. 2.6, different assumptions for that correction lead to widely different results about the homogeneity of the galaxy distribution. While the results in the
left panel of the figure seem more sensible (after all, one should use cosmological
distances and K-corrections), Joyce et al. (1999) have shown that the main factor at these distances is the form of the K-correction used and it is probably not
03
01
76
2002 by CRC Press LLC
63
Figure 2.6 The number-redshift relation for a deep (ESP) redshift catalog for different
;
)
volume-limited subsamples, described in the legend. Cosmological ( M
distances and the K-correction are used for the upper panel; Euclidean distances and no
K-correction are used for the lower panel. The numbers at the straight lines show the
exponents n for the number-distance relation N Rn . (Reproduced, with permission,
from Scaramella et al. 1998, Astr. Astrophys., 334, 404408.)
=1
=0
77
2002 by CRC Press LLC
64
correct. Note also that the formulae of cosmological distances are based on the
assumption of homogeneity that we are checking for.
It is possible to use another form of number counts, up to a limiting apparent
magnitude. This method does not require difficult redshift measurements, and it
was used already by Hubble in 1926.
We know that the observed flux density S from a galaxy of intrinsic luminosity
L at a distance r is (in a Euclidean universe)
S
L
= 4r
2:
(2.35)
All other galaxies of luminosity L and closer than r will look brighter than S , and
their number N S 0 > S is
3=2
nL
0
p LS
N S > S jL
nLV r
;
) = ( ) ( ) = (4 )
( )
= 2:5log10 S + const
N m0 < m
) / 100:6m :
Astronomers usually use differential number counts. These have to follow the
same power law (with a different normalization):
( ) 10
dN m
/ 0:6m :
dm
In Fig. 2.7 we show the number countapparent magnitude relation in the nearinfrared region of the spectra (McCracken et al. 2000). For smaller magnitudes
the classical 2/3 slope is well seen, but for the faint end (larger apparent magnitudes) the slope has to be modeled. These surveys extend to rather large redshifts
(z 12) where the cosmological effects become important. In addition to the
different volume-redshift dependence and the K-correction we also must consider
the evolution of galaxies, which changes the galaxy luminosity function.
Thus, homogeneity of the spatial distribution of galaxies is not easy to establish observationally. It is difficult to disprove it, as well, because all the present
cosmological machinery is based on the cosmological principle, and discarding
a part of it would force one to rebuild the entire edifice from the start. Thus the
78
2002 by CRC Press LLC
65
Figure 2.7 Compilation of differential galaxy number counts in the K-filter, compared to
model predictions. (Reproduced, with permission, from McCracken et al. 2000, Mon. Not.
R. Astr. Soc., 311, 707718. Blackwell Science Ltd.)
main argument for homogeneity is indirect, adding up bit by bit from the success
of all other achievements in cosmology.
H 0 = 65 km/s/Mpc
66
Figure 2.8 Expected abundances of light elements for different values of the mean baryon
density; the vertical shaded region shows the values of the density that are compatible with
observed abundances. (Reproduced, with permission, from Schramm and Turner 1998, Rev.
Mod. Phys., 70, 303318. Copyright 1998 by the American Physical Society.)
This is also one of the few cases in cosmology where we know the essential
parameters of the cosmological model. In the early epochs the energy density of
the matter and the role of the cosmological constant were both negligible compared to the energy density of radiation, and that is well known, determined by
the present temperature of the CMB radiation.
The yields of light elements, except that of helium, depend strongly on the
mean baryon density B in the universe and comparison with observations allows us to pinpoint it well, as shown in Fig. 2.8 (Schramm and Turner 1998). The
main source of uncertainty is in estimating how primordial the observed abundances are. In general, the calculated abundances are in good agreement with
observations, and thus support the overall picture of standard cosmology.
The inferred baryon density B is about twice that for luminous matter, which
is natural, because, for example, there is a lot of mass in intra-cluster gas. An
80
2002 by CRC Press LLC
67
important point is that this parameter is much less than the estimates we have for
the total M 0.21.0, showing that the bulk of the matter (dark matter) must
be nonbaryonic.
= M + 5log10 (DL (
M ;
; H0 ; z)=1 Mpc) + K (z) + AG + 25; (2.36)
where the luminosity distance DL (z ) can be computed using the formulae (2.32,
m
2.29, 2.27 and 2.21) above. The galactic extinction accounts for the absorption of
light in our Galaxy.
If we can find a class of objects with the same luminosity (the same absolute magnitude M , standard candles), then measuring the apparent magnitudes
for these candles for different redshifts, we can fit the observed m z relation
to the formula (2.36) and find the parameters. This has been tried before, using
high-luminosity galaxies in clusters and even quasars, but it was realized quickly
that these objects evolve rapidly, and the m z relation can tell us more about
their luminosity evolution than about the cosmological model. The redshifts observed then were small, too, and the DL z relation does not depend much on the
cosmological parameters for z .
This method has been resurrected, using a new class of standard candles
distant supernovae. Supernovae are stellar explosions that do not depend on their
environment, as galaxies do. They are also very bright and can be observed at
high redshifts, but their main advantage is that their luminosity changes with the
time-scales of weeks to months, allowing better selection of standard supernovae
and better estimation of their intrinsic luminosities.
There are two groups currently observing distant supernovae: The Supernova
Cosmology Project, led by S. Perlmutter (Lawrence Berkeley Laboratory, U.S.A.)
and the High-Z Supernova Search Team, led by B. Schmidt (Mt. Stromlo and
Siding Spring Observatories, Australia). Both groups are highly international and
have discovered about 50 distant supernovae to date; that number is, however, less
than the total membership of both groups. As you can see from the references, the
number of authors of observational papers in cosmology is large, reflecting the
()
()
()
81
2002 by CRC Press LLC
68
Figure 2.9 The dependence of the apparent magnitude mB of distant supernovae on their
redshift z . (Reproduced, with permission, from Perlmutter et al. 1999, Astrophys. J., 516,
565586. AAS.)
fact that the observations are very difficult undertakings that require the cooperation of many observatories.
Fig. 2.9 shows the typical redshift-magnitude relation (for 42 distant supernovae) from the first group (Perlmutter et al. 1999). Different curves show the expected relation for different cosmologies, and we see that this difference is rather
. The best parameter estimate obtained from supernovae
small even up to z
km sec 1 Mpc 1 . It is not easy
thus far is that of the Hubble constant H0
to get good individual estimates of the remaining parameters from the magnituderedshift relation, because both M and influence the luminosity distance in
similar ways and thus the estimates are highly correlated.
Another observationally very active branch of cosmology at the moment is
measurement of the fluctuation spectra of the CMB radiation. Exact modeling
of the recombination process predicts that for certain characteristic angular distances these fluctuations have an enhanced amplitude, and these distances and the
amplitudes (oscillations) depend strongly on the cosmological model. Fig. 2.10
shows a compilation of observational data with a best-fit model (Hu 2001). Most
of the data points in this figure correspond to a whole experiment.
This fit also allows us to estimate the cosmological parameters, and the estimates it gives are also well correlated. A good point is that the correlations are different from those arising from the m z test, and combining the two tests can give
us for the first time a rather well-defined estimate of the parameters. There have
=1
= 65 2
()
82
2002 by CRC Press LLC
69
Figure 2.10 Compilation of the CMB temperature fluctuation amplitudes T=T and the
effective multipole values l for recent experiments (courtesy of W. Hu). The gray boxes
show the bandwidthrms error boxes for different experiments; the light line shows the
best model fit.
been several such determinations recently with similar results. Fig. 2.11 shows
the confidence regions for the parameters from the Boomerang measurements of
the CMB anisotropy and supernovae data fits.
As we see, the different methods limit the possible region of parameters efmodels that are
fectively, and this region also includes the line of the flat k
preferred by theories of the extremely early universe (inflationary models). We
also see that the observations definitely require a cosmological constant. Up to
now many cosmologists have tried to ignore this need because there is not much
physical reason to include the cosmological term in the Einstein equations.
Another problem is the explanation of the numerical values of the parameters.
Because these values have an extremely different behavior in time, the fact that
they have comparable values at the present moment is rather mysterious. One
possibility is to introduce a new physical field that mimics the role of the cosmological constant in the Friedmann equations, but changes slowly in time and can
thus more easily have M . To the four main constituents of the universe
(radiation, baryonic matter, and hot and cold dark matter), we add the new, yet
completely hypothetical fifth field that has been termed quintessence.
=0
83
2002 by CRC Press LLC
70
()
=0
Although the estimates of the parameters described above are among the best
at the moment, they should be used with caution. Cosmological observations are
difficult. Interpretation relies on a rather advanced theoretical machinery, and the
more complex a machine is, the more easily it can break down. We should bear
in mind a well-known fact from the 70-year history of contemporary cosmology
the best estimate of Hubbles constant has changed during these years from
480 km sec 1 Mpc 1 , as found by E. Hubble in 1929, to 65 km sec 1 Mpc 1 , as
determined from the recent observations of distant supernovae. This means that
our mental picture of the universe, our model universe, is almost ten times larger
and older now than it was 70 years ago.
84
2002 by CRC Press LLC
CHAPTER 3
72
1. Local finiteness Any bounded set of IRd contains a finite number of points
of .
2. Simplicity There are no multiple points.
The term point process and the symbol are used with a double meaning;
in fact, they are two different aspects of the same well-defined mathematical concept:
(i) The point process is considered a random sequence of discrete points, =
fx1 ; x2 ; : : :g, so the notation x 2 emphasizes that point x belongs to the
point process.
(ii) The point process is a random measure counting the number of points lying in
a given region A IRd , (A) = n.
The number of points lying in A can be expressed as
(A) =
x2
1lA (x);
1
0
for x 2 A;
for x 2
= A:
(3.1)
We shall use the terms point processes and point fields interchangeably. The
term process was originally introduced because it was associated with time series. In our context, however, it would be more appropriate to use the term field
as Stoyan and Stoyan (1994) have proposed, but we shall also retain the classical
term process. We shall typically deal with point fields in IR3 , and the bounded
region where a given realization of the point field lies will be called the window
and will be denoted by W . The volume of a region A IR3 will be denoted by
V (A) and the probability that k points of the process lie in A by P ((A) = k).
Although the expressions in the following sections are given for point fields in
the Euclidean space IR3 , it is important to keep in mind what was stated in Chapter
2 regarding the geometry of the space depending on the adopted cosmological
model. Distances and volumes have to be calculated according to such a model.
86
2002 by CRC Press LLC
Point processes
73
(x) = lim
dV !0
h(dV )i ;
(3.2)
dV
where the bracket h:::i denotes expected value of the random variable. For simplicity, we represent the infinitesimal region and its volume with the same notation, dV .
A point process is called stationary or homogeneous if its statistical properties
are invariant under translations; if the invariance holds under rotations the process
is called isotropic, and in that case there is no privileged direction.
For a homogeneous process the intensity function takes a constant value ( ) =
, which is the mean number of points per unit volume. This value is called num.
ber density in cosmology, and it is often denoted by n
We can similarly define the second-order intensity function
2 (x1 ; x2 ) =
lim
dV1 ;dV2 !0
h(dV1 )(dV2 )i :
dV1 dV2
(3.3)
Again, if the process is isotropic and homogeneous this quantity depends only on
j=r
the distance j
x y
2 (r) = 2 (x; y) :
=
N
:
V (W )
h(A)i = V (A):
The name of the process is quite natural if we consider that the form of the probability function of finding exactly k points in A is just
P ((A) = k) =
N
k
pkA (1 pA )N k ;
87
k = 0; 1; 2; : : : ; N;
74
where
pA =
V (A)
:
V (W )
This kind of process is very often used in cosmology; in particular, auxiliary random catalogs, which are binomial fields, are used in the estimation of the twopoint correlation function (see Section 3.4.1). A binomial random field is typically
referred to as a Poisson catalog, random distribution, or Monte Carlo sample with
N points. There is, however, a subtle difference between a homogeneous Poisson
field and a binomial field. Strictly speaking, as we shall see in the next section, in
a homogeneous Poisson field the number of points is also random.
3.2.3 Poisson processes
A fundamental property of a Poisson field is that if we consider a finite number
k of disjoint regions Ai , the random variables f(Ai )gki=1 , providing the number
of points of the process lying in Ai , are stochastically independent (Stoyan and
Stoyan 1994). We can distinguish two kinds of Poisson processes, called, respectively, homogeneous or stationary and inhomogeneous or general.
Homogeneous Poisson processes
In this kind of process ( ) does not depend on position and therefore has a
constant value ( ) = .
The name of the process is quite natural if we take into account that the number
of events of the point process lying in a region A of space with volume V (A) is a
random variable, (A), having a Poisson distribution with parameter V (A),
P ((A) = k) =
[V (A)]k
e
k!
V (A) ;
k = 0; 1; : : :
88
2002 by CRC Press LLC
75
P ((A) = k) =
[(A)]k (A)
e
;
k!
k = 0; 1; : : : ;
(A) =
(x)dx:
It is clear from these formulae that, in general, the stationarity property is lost.
For example, in Fig. 3.1 we show a simulation of an inhomogeneous Poisson
process for which the intensity function follows a Plummers model, a polytrope
of index 5:
(x) /
p
x2
1+ 2
R
5
2
where x = x21 + x22 + x23 . The density decreases radially from the center of the
cube following the curve showed in the bottom panel of Fig. 3.1.
In Chapter 5 we will study several models of point fields that have been used
in different contexts of the description of the large-scale distribution of matter in
the universe.
3.3 The relation between discrete and continuous distributions
As we have stated in the introduction to this chapter, the distribution of mass in
the universe can be regarded as a continuous density field. This will be formalized
in Chapter 7. The luminosity, however, is concentrated into galaxies, forming a
discrete point process. In this section we discuss the connection between these
two approaches, in particular how the density field can be estimated from the
galaxy distribution.
The three-dimensional distribution of galaxies can be regarded as a point field,
where f i gN
i=1 represents the coordinates of the positions. One way, although not
the only one, to establish a connection between the discrete galaxy distribution
and the underlying continuous density field is by means of an inhomogeneous
Poisson point process with density n( ) described formally as a sum of Dirac
delta functions D ( ):
n(x) =
N
X
i=1
mi D (x
x
x
xi )
(3.4)
89
2002 by CRC Press LLC
76
4500
4000
3500
(x)
3000
2500
2000
1500
1000
500
0
0
0.2
0.4
0.6
0.8
Figure 3.1 Top: A 3-D simulation of an inhomogeneous Poisson process having a polytrope of index 5 as intensity function with R = 0:5; the center of the box is the origin of
coordinates and the sidelength of the cube is 1. There are 1000 points in the simulation.
Bottom: The intensity function of that process.
90
2002 by CRC Press LLC
77
(x) = n [1 + (x)] :
Here n
= E f( )g is the expectation value of the intensity over all realizations
of the density field. We have assumed that density is a homogeneous random
can be estimated by the mean number density over the
field; then by ergodicity n
whole sample volume. The random field ( ) is the density contrast (a zero-mean
dimensionless field).
Now, if we place randomly a volume dV at position , the probability that it
contains a point is
P = (x)dV:
This provides the basis of what is called in cosmology the Poisson model, introduced by Layzer (1956) (see also Peebles 1980, x 33 and Fry 1985 for a more
detailed presentation). Using the statistical terminology, this is a double-stochastic
point process, because there are two different levels of randomness: the random
field itself and the Poisson sampling.
3.3.1 Estimators of the density field: intensity functions
If galaxies are considered Poisson tracers of a continuous density field, we could
be interested in estimating the underlying density field from the point field. The
usual way of doing this is to convolve the spatial galaxy distribution with a window function in order to get a smooth field. Of course, under this hypothesis we
are considering that galaxies trace mass, but this may not necessarily be true,
and different biasing schemes can be adopted, once the density field has been
estimated by considering that the galaxy formation probability is not just proportional to the local density, but instead follows more complicated dependence of
the density environment (Coles 1993; Dekel and Lahav 1999). Biasing schemes
are introduced in Section 7.9.
When cosmologists estimate the density field from the discrete galaxy distribution, they are making a kernel estimation of the intensity function (Saunders et al.
1991).
Let f 1 ; 2 ; : : : ; N g be the positions of N galaxies in a bounded region W .
The estimator of the intensity function with kernel ! (also called filter function)
and smoothing radius (bandwidth) ! > 0 is
x x
b! (x) =
N
X
i=1
! (x
xi); x 2 W;
(3.5)
where ! is a symmetric density probability function. In the field of point processes the selection of the kernel function is not considered as relevant as the
This process is also referred to as a Cox process. A more complete treatment is given in Chapter 5.
91
2002 by CRC Press LLC
78
1
0.8
1.5
0.6
0.4
0.5
0.2
0
-1.5
1.5
1
0.5
-1
-0.5
0.5
0
0
-0.5
0.5
-1
-0.5
-0.5
-1
1.5 -1.5
0.5
1 -1
Figure 3.2 The Gaussian kernel (left) and the Epanechnikov kernel (right) for the same
value of the bandwidth, ! = 0:5.
choice of the value for the smoothing radius ! . The reconstructed density field
strongly depends on the adopted value for ! .
The standard kernel function used in cosmology is the Gaussian filter
! (y) =
1
exp
(2)3=2 !3
jyj2 :
2!2
k! (x) =
x2
3
4! 1
!2
for j j !
otherwise
(3.6)
In Fig. 3.2 we show a two-dimensional representation of these two kernel functions, and in Fig. 3.3 we show a planar cluster point process and the kernel
smoothed density field reconstructed by means of a Gaussian filter with different bandwidths.
An optimal selection of ! when one wants to estimate the intensity function
from a point process is a well-studied problem in the mainstream of spatial statistics (Diggle 1983; Diggle 1985; Silverman 1986; Wand and Jones 1995) and it is a
clear example of how the collaboration between astronomers and statisticians can
be useful. For example, using adaptive intensity estimators would help in regions
where the density varies quite abruptly (Stein 1997). A simple generalization of
the Gaussian kernel with this goal is
A (y) =
where
(2)3=2 det(A)1=2
exp
1 T
y Ay :
2
79
=0.04
=0.02
=0.08
Figure 3.3 A clustered planar point field and its reconstructed density field by means of a
Gaussian filter with three values of the smoothing radius.
x x
(3.7)
x x
93
2002 by CRC Press LLC
80
r) of a homogeneous
(3.8)
x2)(x1):
(3.9)
The first term comes from the statistical independence of the process in separate
volumes and the second term comes from self-pairs. It can be derived, dividing
the volumes dV1 ; dV2 into very small cells, so that the occupation numbers ni of
these cells are either 0 or 1 (Peebles 1980, x 36). Then the average
X
hn2 (x)i dV = h
n2i i dV = (x)dV:
x x
x2)n:
(3.10)
Omitting the last term, as we have explained above, and comparing (3.10) with
the formula for the correlation function (3.8), we get
(r) = R (r);
x x
x x
(r)
(r) = 2 2
n
1;
(3.11)
although in other fields of physics and statistics, the pair correlation function,
also called the radial distribution function or the structure function, is just g (r) =
(r) + 1.
94
2002 by CRC Press LLC
81
bmin (r) =
Nin
V (W ) X
ni (r)
NNin i=1 Vsh
1;
(3.12)
where the quantity bmin (r) is the minus-estimator of the correlation function and
Vsh is the volume of the shell of width dr,
4
Vsh = [(r + dr)3 r3 ];
3
which can be approximated by 4r2 dr if dr is small, whereas ni (r) is the number
of points of the process within the whole window W lying at a distance between
r and r + dr from point i, which lies in Win . As we can see in Fig. 3.4, the scale
r + dr appearing in (3.12) has to be less than or equal to rmax , the distance between the boundary of the inner window Win and the edge of W . One can then
choose between two possibilities: to fix rmax as the maximum scale at which we
want to compute (r), or to shrink the window Win as r increases, in such a way
that rmax varies, being for a given scale rmax = r + dr. The advantage of this
estimator is that it can be applied safely to nonhomogeneous point fields, because
no assumptions are made about the distribution of points outside W . One drawback is that this estimator does eliminate some of the information contained in the
data. Moreover, at large distances only a small fraction of the galaxies are considered as centers, increasing the variance of the estimator (Kerscher, Szapudi, and
Szalay 2000). If one wants to make full use of the data contained in the catalog,
an edge correction has to be applied.
Several edge-corrected estimators of are commonly used in cosmology. The
simplest edge correction is incorporated in the so-called natural estimator, first
used for the study of the angular correlation function by Peebles and Hauser
(1974). An auxiliary random sample containing Nrd points must be generated
in W (i.e., a binomial point process with Nrd points), and the estimator for the
correlation function is defined as
bPH
Nrd 2 DD(r)
=
N
RR(r)
1;
(3.13)
where DD(r) is the number of pairs in the catalog (within the window W ) inside
95
2002 by CRC Press LLC
82
Figure 3.4 The inner window used in the minus-estimator of the correlation function. Only
points within Win are considered as centers for counting neighbors.
the interval [r; r + dr], and RR(r) is the number of pairs in the random catalog
with separation in the interval mentioned above. The use of random samples in
this and other estimators replaces the calculation of partial shell volumes; it is,
in fact, the Monte Carlo integration for volumes. This also means that the number of random points must be much larger than the size of the data sample; ten
times is usually considered sufficient, but more could be needed for the DR pairs
(see below). The normalization factor in Eq. 3.13 (Nrd =N )2 can be replaced by
(Nrd (Nrd 1))=(N (N 1)). In this case, the estimator is unbiased for a Poisson
process. This choice is made by some authors and applies to all other estimators
below that include this normalization constant.
However, the natural estimator has been shown to suffer from insufficient correction for edge effects. A better and extensively used estimator is that of Davis
and Peebles (1983), sometimes called the standard estimator:
N DD(r)
bDP (r) = rd
N DR(r)
1;
(3.14)
where DR(r) is the number of pairs between the data and the binomial random
sample with separation in the same interval.
Hamilton (1993b) calculated the bias of the above two estimators and found
that the expectation value for the standard estimator (3.14) can be written as
E fbDP (r)g =
where b(r) is the unbiased estimator of the correlation function, the mean over , the
density measures the error we make when we use for the mean density, n
(1 + ), and (r) is the galaxy
estimate nest found from our sample, nest = n
96
2002 by CRC Press LLC
83
bHAM (r) =
DD(r) RR(r)
[DR(r)]2
1;
(3.15)
which has only a second-order bias, caused by the finite sample effects:
E fbHAM (r)g =
b(r) 2 (r)
:
[1 + (r)]2
The DR(r) term in the above estimator (3.15) may introduce numerical noise
at small distances. To eliminate that we must either choose larger random samples (Pons-Bordera et al. 1999) or integrate the angular part of the pair integral
analytically, as Hamilton (1993b) does.
Another estimator, proposed almost simultaneously by Landy and Szalay (1993),
has similar properties:
bLS (r) = 1 +
Nrd 2 DD(r)
N
RR(r)
N DR(r)
2 rd
:
N RR(r)
(3.16)
The relation between the natural and LandySzalay estimators can be easily deduced from their definitions given in Eqs. 3.13 and 3.16:
N DR(r)
bLS = bPH + 2 2 rd
:
N RR(r)
(3.17)
The LandySzalay estimator has also a second-order bias, if the mean density
is defined from the sample data, as in (3.16) above:
E fbLS (r)g =
b(r) 2 (r) + 2
:
[(1 + ]2
(3.18)
The authors of this estimator, however, propose to estimate the mean density by
independent means (e.g., by a maximum likelihood method). Then we must substitute the mean overdensity in (3.18) with the fractional error in the mean density
, which could be much less than , further reducing the bias.
Recently, Szapudi and Szalay (1998) generalized the LS-estimator for higherorder correlation functions. They argue that it is the most natural estimator; if we
97
2002 by CRC Press LLC
84
Figure 3.5 In this illustration we see the volumes entering the definition of the Rivolo
estimator for (r ). If the shell touches the boundary of the window, only its intersection
with W is considered.
bPH (r) =
brackets hi denote sample averages, we see that it is unbiased (for infinite samples), but the linear terms in (3.19) will generate additional terms in the sample
variance of . The symbolic presentation for the LS-estimator is
bLS (r) =
(De 1
Re)(De 2 Re)
= h1 2 i;
g
RR
which has only the necessary term for the correlation function, leading to an efficient (minimum-variance) estimator. This can be proved exactly; we shall describe that below.
If we have simple sample volumes, we can find the partial shell volumes without using the Monte Carlo trick of comparison with random samples. The simplest
estimator of that kind is that of Rivolo (1986):
bRIV (r)
N
ni (r)
V (W ) X
=
N 2 i=1 Vi (r)
1;
(3.20)
where ni (r) is the number of neighbors at distance in the interval [r; r + dr] from
galaxy i and Vi (r) is the volume of the intersection with W of the shell centered
at the ith galaxy and with radii r and r + dr. This is illustrated in 2-D in Fig. 3.5.
98
2002 by CRC Press LLC
85
Figure 3.6 A 2-D illustration of the denominator in the OS estimator of the correlation
function, Wx Wy . (Reproduced, with permission, from Pons-Bordera et al. 1999, Astrophys. J., 523, 480491. AAS.)
bOS (r) =
n X
n
k(r jxi
V (W )2 X
2
2
N 4r i=1 =1 V (W \ Wx
j
j =i
xj j)
i
xj )
1;
(3.21)
where k (x) is a kernel smoothing function (see also Fiksel 1988). The Epanechnikov kernel (3.6) is typically used in point field statistics.
Here Wy denotes the window W shifted by the vector , Wy = W + = f :
= + ; 2 W g. The denominator is the volume of the window intersected
with a version of the window that has been shifted by the vector i
j . It also
can be written as Wxi \ Wxj (see Fig. 3.6). The set covariance function is defined
as
W ( ) = V (W \ Wx ) (Stoyan, Kendall, and Mecke 1995).
We can still improve the OS estimator by replacing 4r2 in the denominator
2
of (3.21) by the quantity 4 j i
j j . In particular, Stoyan and Stoyan (2000)
recommend this replacement for small r and large value of the bandwidth ! of the
kernel function. They also have shown that a geometric version of the Hamilton
estimator can be obtained by improving the density estimator entering in (3.21).
In fact, the square of the intensity is just estimated as
x z yz
x
x x
2 =
N 2
;
V (W )
99
y
x x
86
while the improved estimators for the intensity have the general form
ep (r) =
p(xi ; r)
;
C (r)
i=1
N
X
where p(
V (W \ b(x; r))
:
4 r3
3
pV (x; r) =
(3.23)
These two functions allow us to define the corresponding weighted improved ineS (r) and
eV (r).
tensity estimators
The Stoyan estimator for the correlation function is then
bSTO (r) =
n X
n
X
k(r jxi
1
2
e
V
S (r)4r i=1 =1 (W \ Wx
j
j =i
xj j)
i
xj )
1:
(3.24)
Stoyan and Stoyan (2000) have derived an approximation formula for the variance
of the estimator valid for Poisson processes. On the basis of this formula and also
considering simulations, they recommend the use of the rectangular kernel
k!r (r) =
1
1l
2! [
!;!] (r )
87
10
(r)
DP
HAM
1
LS
RIV
OS
PH
0.1
1
10
r (h-1 Mpc)
Figure 3.7 The correlation function of a volume-limited sample of the PerseusPisces catalog calculated by means of six estimators described in the text.
As we can see in Fig. 3.7, there is not much difference between the estimates
of (r) provided by the formulae above when applied to the sample. We should
mention only the departure of the OS estimator at small separations, taking higher
values than the others, probably as a consequence of its lesser sensitivity to local
anisotropies due to peculiar motions.
Several attempts have been made to compare different estimators, using real
data samples, N -body simulations, and point processes with known correlation
functions (Pons-Bordera et al. 1999, Kerscher, Szapudi, and Szalay 2000; Stoyan
and Stoyan 2000). The results are the same; the differences between the estimates
are more relevant at large scales where (r) 1 and fluctuations in the mean
density affect the estimators more strongly. Pons-Bordera et al. (1999) find that
at large scales the Hamilton and LandySzalay estimators provide the best results.
Kerscher, Szapudi, and Szalay (2000) recommend the LandySzalay estimator
because it is easier than the Hamilton estimator to calculate numerically.
Errors and variances in (r)
The theoretical formula for the variance of the two-point correlation function depends on the three- and four-point correlation function (Peebles 1973, Hamilton
1993b). We shall follow Hamilton (1993b), and write the unbiased estimator b(r)
101
2002 by CRC Press LLC
88
b(r12 ) = 12 =
hW12 1 2 i ;
hW12 i
(3.25)
r r
where the window function W12 selects the proper coordinate shift 1
2 = 12 ,
2
(
;
+
)
,
and
the
angle
brackets
denote
being nonzero for 1
2
12
12
sample averages:
r r
dr r
hW12 1 2 i = V12
Z Z
dr
dr
The finite width 2j j of the window function represents the usual way we measure the correlation function, binning it into separate intervals; this can also be
thought of as averaging over that interval. The window function can be written as
W12 = w12 1 2 ;
(3.26)
where w12 is the pairwise weighting function, and i is the catalog selection
function at the point i .
The covariance of the estimator (3.25) for two different coordinate intervals 12
and 34 is
34
34
(3.27)
e(r) =
f (r)b(r)dV;
hh2 eii =
ZZZZ
f12 W
f34 (13 24 + 14 23 + 1234 ) dV1 dV2 dV3 dV4 ;
f12 f34 W
(3.28)
where 12 (r12 ) is the true (population) two-point correlation function, 1234
f denotes normalized pair
is the population four-point correlation function, and W
windows,
f12 =
W
W12
:
W12 dV1
89
There are -function terms in the integral (3.28), where some of the points 14
coincide. Separating these, we arrive finally at
Var e =
ZZZZ
f12 f34 we34 3 4 (1234 + 13 24 + 14 23 ) dV1 dV2 dV3 dV4
ZZZ
4
(3.29)
f f we ( + ) dV dV dV
n Z Z 12 13 13 3 123 23 1 2 3
2
2 we (1 + ) dV dV ;
+ 2
f12
12
12
1 2
n
where n
is the mean number density and 123 is the three-point correlation func+
tion of the population. The first member in the sum (3.29) results from disjoint
pairs 12 and 34, the second member comes from the configurations where the
pairs share one point, and the last term comes from the case where the two pairs
coincide. As we see, this variance includes both four- and three-point correlation
functions, and is thus rather difficult to calculate.
The variance (3.29) depends on the pair weight function. Minimizing it in respect to the pair weighting gives an approximate result for the best weight function
(Hamilton 1993b):
w12
1
;
[1 + 4n J3 (r)]2
where is the selection function at the location of the pair and J3 (r) is the integral
of the correlation function (Peebles 1980):
1
J3 (r) =
4
(r)dV =
Z r
(s)s2 ds:
= wi wj is used, with
1
wi =
:
1 + 4n i J3 (r)
(3.30)
(3.31)
As the integral J3 (r) depends on the pair separation, different weights should be
chosen for different separations. This weighting is similar to the FKP weighting
used in determination of power spectra (see Section 8.2). For dense regions, the
weighting is inversely proportional to the local density and volume elements are
weighted equally; for sparsely sampled regions, the weight function approaches
unity, giving equal weight to separate galaxies.
J3 (r) in (3.30) gives the average number of neighbors in exThe quantity 4 n
cess over a random distribution up to a distance r for any sample point. It is usually determined from theoretical considerations. If we try to determine it on the
basis of a point sample of size N , the value of that quantity for the whole sample
volume will be 1, as every point has then N 1 neighbors, one less than expected for a random distribution (Peebles 1980, x 32). This integral constraint will
J3 (r) from
cause the decrease of J3 (r) at large separations. If we determine 4 n
103
2002 by CRC Press LLC
90
our point sample, we should stop when it reaches the maximum at a separation
rmax and use this maximum value for larger separations r rmax .
Apart from the variance caused by the correlation structure of the data, the
other sources of variance are the finite volume of the sample, insufficient edge
correction, and the discreteness noise. These errors are usually coupled together
as cosmic variance (Szapudi and Colombi 1996), although the term is more
appropriate for the errors caused by the finite size of the sample. We have seen
above how the different finite volume effects affect the bias of the estimator. These
terms also contribute also to the variance; the smaller the volume of our sample,
the larger the difference between our estimate of the correlation function and the
actual correlation function.
In the case of a homogeneous point process, if we sample a large enough volume, we expect to obtain statistically reliable results, which approximate well
the correlation function of the entire universe of galaxies. This is the fair sample hypothesis (Peebles 1980), which encompasses the cosmological principle,
and the ergodic hypothesis introduced in Section 1.3. The size of a fair sample
and the cosmic error that results from the finite size of the sample are difficult to
estimate. Because these estimates require information about scales that have not
yet been directly observed, we can find these estimates only by using theoretical
assumptions.
The variance that is caused by insufficient edge corrections and by the discreteness noise is easier to estimate, in principle, especially for certain point processes.
Ripley (1988) has discussed extensively the variance of second-order estimators
for Poisson and binomial processes. The variance of these estimators contains
2 and the much larger O(n 1 )
both the Poisson pair term that is proportional to n
terms. The latter terms are caused by insufficient correction for edge effects. Analytical expressions of the intrinsic variances of several estimators are also given
in Hamilton (1993b) and Landy and Szalay (1993).
As noted above, Szapudi and Szalay (1998) proposed the L-S type estimators
for correlation functions of any order. Their estimators can be written symbolically as
bN =
1X N
( 1)N
S i i
D
i
R
N i
(3.32)
where and are the intensities of data and the random reference process, respectively. Each factor in the sum (3.32) has to be calculated at a different point
(N -point correlation functions depend on N arguments). Note that the overall
104
2002 by CRC Press LLC
91
S=
(x)N ;
where N is the N -dimensional Lebesgue measure and (x1 ; : : : ; xN ) is a function that selects a specific configuration of N points (e.g., for the two-point correlation function (x1 ; x2 ) equals one when the distance between the points x1
and x2 falls into a given distance bin, and is zero otherwise).
Using the technique of factorial moment measures, Szapudi (2000) showed that
the average of the estimator is
Z
Var =
1
;
Np
DD (r) =
DD(r):
Since similar expressions are valid for DR(r) and RR(r), it is quite straightforward to obtain an approximate expression of the Poisson errors associated with
the estimators bPH , bDP , bHAM , and bLS :
(r) '
1 + (r)
;
DD(r)
105
2002 by CRC Press LLC
(3.33)
92
where it has been considered that the terms 1=DR(r) and 1=RR(r) vanish when
choosing a large enough random sample.
Alternatively, we can estimate the Poisson errors by using in (3.33) instead
g (r) or
of the number of data pairs DD(r) the number of datarandom pairs DR
g (r), which are normalized to the total
the number of randomrandom pairs RR
number of objects in the catalog:
g (r)
DR
N
DR(r);
=
Nrd
g (r)
RR
N 2
=
RR(r):
Nrd
These pair numbers give a better approximation of the Poisson error, especially
for smaller scales with more data pairs.
For geometric estimators, Poisson errors have similar expressions. For example, for the Rivolo estimator
RIV (r) =
2V (W )(1 + (r))
;
P
N Ni=1 Vi (r)
while for the Stoyan estimator with the rectangular kernel the approximate error
has the form
s
STO (r) =
where
1 + (r )
;
4r2 !e2S (r)
W (r)
1 2
sin()
W (x(r; ; ))dd ;
4 0 0
and
W (r) is the isotropized set covariance function of W .
W (r) =
For cluster processes, the Poisson method underestimates the errors. We can
obtain this result if we use only part (the last term) of the minimum variance expression (3.29). Real estimators have additional errors caused by the edge effects;
these, too, are ignored here. A rough rule is that Poisson errors give an orderof-magnitude estimate of the sample errors. Thus, frequently the errors given by
(3.33) are multiplied by an empirical coefficient in the range 12 (Mo, Jing, and
Borner 1992; Martnez et al. 1993).
2. Poisson enhanced errors. The assumption made previously about the randomness of the point distribution is in general not true for the clustered galaxy
distribution. Peebles (1980) and Kaiser (1986) proposed estimating the variance
of the correlation function on the basis of the cluster model. This model, explained in more detail in Peebles (1980, x 40), assumes that all points are contained in isolated clusters. The mean number of points per cluster can be found
as the mean number of neighbors for any point. The definition of the correlation
function allows to write the probability dP (r) of finding a neighbor for a point at
the distance r in the volume element dV as
(3.34)
93
Nc = n
(r)dV
(r)r2 dr;
where the last equality assumes that the integral exists. Usually correlation functions fall off rapidly enough for that.
The integral J3 has a meaning only in the context of the cluster model, where
the pair separation range in the integral is thought to extend only to a typical
cluster size. In this case 4J3 = Nc 1, where Nc is the number of points in a
cluster, and averaging it over all clusters with different numbers of points gives
(Peebles 1980, x 31)
4n J3 =
For pair separations larger than the typical cluster size the fluctuations in the
correlation function are Poissonian in the number of clusters N 0 = N=Nc , where
N is the total number of galaxies in the sample. This leads to the enhanced
Poisson errors, given by
(r) =
1 + 4n J3
p
:
DD(r)
(3.35)
BOO (r) =
v
uM
uX
t
(3.36)
where i (r) is the two-point correlation function calculated on the bootstrap sample labeled by i and
M
1 X
(r):
(r) =
M i=1 i
94
of a good idea. About one third of points in a generated sample, selected with replacements, are multiple. This is certainly not a good representation of the parent
population, where there are no multiple points. Another possibility is to discard
multiple points; in this case, the generated sample has about two thirds of the
density of the original sample (and the parent population). Thus, there is only one
possibility to generate a sample with the same correlation structure that the original sample. We have to select the same number of points without replacement,
getting back the original sample. Kerscher, Szapudi, and Szalay (2000) express
that idea, saying that in spatial statistics, a whole sample takes the role of one
point in the bootstrap procedure. Moreover, Snethlage (1999) has shown that the
bootstrap variance of the correlation function can be calculated analytically, without resorting to the bootstrap procedure, and that it differs completely from the
real variance.
4. Disjoint subsampling. If the sample volume is large enough, we can measure
the correlation function in different disjoint subregions and take as a measure of
the error the dispersion of for the different subregions (Maddox et al. 1990b,
Buchert and Martnez 1993). The problem with this method is that it can be used
only for large samples and the accuracy of the variances is inversely correlated
with the size of the subregions. Maddox et al. (1990b) determined the (angular)
correlation function of the APM survey at large separations, and used only four
subregions to estimate variances of the correlation function.
5. Subregion fluctuations. This method was proposed by Hamilton (1993b). It
estimates the variance of the correlation function estimator on the basis of the
sample data. As Hamilton argues, it automatically includes all statistical sources
of error.
This method can be derived by considering the correlation function and its
estimator as functions of the pair window, (r) = (Wij (r)), where the indices
ij refer to all pairs of small volume elements. The error of the estimator is then
caused by the difference of the sample windows W and the unknown population
windows Wpop ; assuming that this difference is small, we can write
= (W ) (Wpop ) =
X
ij
(Wij
@
:
Wij;pop )
@Wij pop
(3.37)
In this sum and below we suppose that the indices i and j are distinct. We also
omit the correlation function argument r, when it is the same throughout a formula. The Taylor expansion (3.37) is valid if both windows are consistent (normalized). The normalization condition for sample windows can be expressed requiring that does not depend on rescaling of the window functions:
@
@ (W ) X
=
Wij
= 0:
@
@Wij
ij
(3.38)
This eliminates the first sum in (3.37). Defining the fluctuation of the correlation
108
2002 by CRC Press LLC
95
@
;
ij = Wij
@Wij pop
(3.39)
=
ij
ij :
(3.40)
@
:
@Wij
ij Wij
As Hamilton notes, this approximation can be done for any single pair, but not for
their sum (3.40), where it will give zero by the normalization condition (3.38).
If we assign the fluctuations ij caused by pairs to their volume elements by
i =
1X
ij ;
2 j
we can change the sum over pairs into the sum over volume elements,
=
X
i
i :
= Wi Wj , the fluctuation
1
@
i = Wi
:
2 @Wi
h(r1 )(r2 )i =
XX
ij
kl
This sum is similar to the integral (3.28) above. Choosing the same value for the
pair distance r and replacing summation over pairs by summation over volume
elements we get the expression for the variance
Var b = h( )2 i =
X
ik
i k :
(3.41)
The expressions for i depend on the form of the specific estimator. Hamilton
(1993b) lists these for the standard, Hamilton, and LandySzalay estimators. For
the Hamilton estimator the fluctuation for a volume element is
DD
i = (1 + est ) i
DD
Di R
DR
109
DRi Ri R
+
;
DR RR
96
where Di D is the pair number between the subvolume i and the full sample for
the pair distance interval (r; r + dr). Other pair numbers have a similar meaning.
In practice, the tiny subvolumes are substituted by finite, but small subregions.
For the correlation analysis of the IRAS 2 Jy survey Hamilton (1993a) split the
survey volume into 528 subregions; for estimating the variance of the power spectrum by the same method (see Section 8.5), Hamilton and Tegmark (2000) split
the PSCz survey into 220 subregions.
The difficulty of this method is the integral constraint; the fluctuations sum to
zero. This also leads to the zero sum for the variance. Hamilton (1993b) proposed
to overcome this by first ordering the subvolume pairs in (3.41) by their mutual
separation and by gradually including into the sum pairs with growing separation,
until the sum stops increasing. This gives an underestimate of the true variance,
but Hamilton (1993b) has shown that for real galaxy samples and for scales that
are not close to the sample size the error is small. We note that the fluctuation
analysis has to be carried out separately for all values of the argument r (argument
intervals) of the correlation function (r).
6. Artificial samples. Since ideally we would like to have a large number of independent samples in order to measure the dispersion of the observed correlation
function, one possibility is to construct artificial samples with correlation function
similar to the observed one. The variance of the correlation function can then be
found by comparing the correlation functions of artificial samples.
This can be done by using N -body simulations resembling the real distribution
of galaxies (Fisher et al. 1993) or by means of realizations of stochastic models with known analytical expressions for . For example, Cox processes, which
will be explained in Chapter 5, are suitable for this purpose (Pons-Bordera et al.
1999). The results, however, must be interpreted with caution, because the higherorder correlations of artificial samples may differ from the higher-order correlations of the actual galaxy distribution, substantially changing the variance of the
two-point correlation function. As we have seen above, at least the three- and
four-point correlation functions must be well modeled to estimate the variance of
the two-point correlation function.
Selection function
The selection function can enter the correlation function estimator explicitly, but
usually it is included indirectly, via the random sample. This sample is generated
to have not only the same geometry, but also the same selection function as the
galaxy sample. Thus errors of the selection function generate additional errors in
the correlation function estimator and, consequently, the selection function must
be determined carefully.
The main method of finding the selection function is the maximum likelihood
estimation of the luminosity function, given by an analytic form. This method is
described in Chapter 1. There are also two other popular methods. In the stepwise
110
2002 by CRC Press LLC
97
likelihood method (Efstathiou, Ellis, and Peterson 1988) the luminosity function
is represented as a step function. This method is thus essentially nonparametric.
Saunders et al. (1990) use both methods and estimate the goodness of the parametric result by the 2 test, comparing the parametric and nonparametric luminosity
functions.
Another possibility is to use the maximum likelihood method directly to find
the selection function. This method was proposed by Saunders et al. (1990) and
is similar to the determination of the luminosity function in Section 1.3.2.
The above methods give the selection functions for the space determined by
the distance measure; it is usually redshift space. If we want to estimate the real
space correlation function, we also have to transform the selection function to real
space. This is rather complex; for details, see Hamilton (1998).
3.4.2 The angular two-point correlation function
Similar to the definition of the correlation function given in Eq. 3.34, we can define the (2-D) angular correlation function w() in such a way that the conditional
probability of finding an object in the solid angle d
at angular separation of an
arbitrarily chosen galaxy is
dP = N [1 + w()]d
;
(3.42)
wbDP () = F
DD()
DR()
1;
(3.43)
where DD() is the number of galaxygalaxy pairs in the catalog with separations in the interval [ d=2; + d=2], and DR() is the number of pairs
between the data and the random sample with separation in the same range; F
is the quotient of the density in the random catalog and the density in the galaxy
catalog.
Angular correlation functions were measured on projected galaxy catalogs much
earlier then the spatial correlation function on redshift surveys (Peebles and Hauser
1974; Groth and Peebles 1977). Recent results on the angular two-point correlation function from the observed sky maps were obtained by Maddox et al. (1996)
for the APM survey described in Section 1.4.1. In Fig. 3.8, we can see these rebDP () for 185
sults for the estimator (3.43). The filled circles show the mean of w
Schmidt fields. In this case the density used to normalize the estimate has been
calculated using the counts in each individual plate. When a global estimate of
the mean surface density of galaxies is used, the estimated value of the correlation function increases at larger angular separations (see open circles in Fig. 3.8).
The reason for this bias when estimating w() is the integral constraint, which
111
2002 by CRC Press LLC
98
Figure 3.8 The angular correlation function of the APM catalog calculated by the DP
estimator (solid circles). At large angular scales the estimate is low biased due to the
integral constraint. The open circles and the stars represent, respectively, the correction
for that bias performed by using a global density in the estimator or by adding a certain
constant that accounts for the power lacking at large scales. (Reproduced, with permission,
from Maddox et al. 1996, Mon. Not. R. Astr. Soc., 283, 12271263. Blackwell Science Ltd.)
will be explained in Section 8.2. In Fig. 3.8, Maddox et al. (1996) show how
bDP makes it possible to correct for the bias.
adding a constant to the estimate w
This constant is estimated from the variance in the field number counts.
3.4.3 The correlation integral
The expected number of points within a distance
galaxy is
hN ir =
Z r
4ns2 (1 + (s))ds =
4 r 2
s 2 (s)ds:
n 0
(3.44)
The last expression may also be referred to as the correlation integral C (r).
K (r) = C (r)=n is called the Ripleys K -function and is used extensively in the
112
2002 by CRC Press LLC
99
K (r) =
Z r
4s2 (1 + (s))ds:
(3.45)
Its main advantage is that no binning of the data is necessary for the estimation.
For a homogeneous Poisson process this function is just
KPois (r) =
4 3
r:
3
L(r) =
(3.46)
L-function
(Ripley 1981)
3K (r) 1=3
4
(r) = nK (r) ;
KPois (r)
(3.47)
100
K (r) is a well-known quantity in the field of spatial statistics and several analytical results regarding its shape and variance are already available for a variety of point processes.
3. It is very important to estimate the quantity K (r) directly from the data and not
through numerical integration of 1 + (r), which introduces artificial smoothing of the results. Several edge-corrected unbiased estimators are available for
K (r). In the context of the present application, the most desirable properties
an estimator must have are to have little variance and not to introduce spurious homogeneity by means of the edge correction. In the next subsection we
comment on different estimators for K (r).
2.
Estimators
We shall make the assumption that the process under consideration is homogeneous and isotropic.
Several estimators exist for K (r). A comparison of some of them can be found
in Doguwa and Upton (1989). From the definition of K and ignoring the edge
effects one could consider the following naive estimator
Kb N (r) =
N X
N
V X
(r
N 2 i=1 j =1
j=
6 i
jxi xj j);
(3.48)
where is Heavisides step function, whose value is 1 when the argument is positive and 0 otherwise. Obviously, for a finite sample this estimator will provide
values for K smaller than the true values since neighbors outside the boundaries
are not considered. One possible solution is to use the counterpart for an integral
quantity of the minus-estimator for (r): Let us consider only points in an inner
region as centers of the balls for counting neighbors. The points lying in the outer
region, a buffer zone (Upton and Fingleton 1985; Buchert and Martnez 1993),
take part in the estimator just as points that could be seen as neighbors at a given
distance r of the points in the inner region. The inner region might shrink as r
increases. However, this solution leads to biases (the sample is not uniformly selected), wastes a lot of data, and obviously increases the variance of the estimator
(Doguwa and Upton 1989). The standard solution adopted in the statistical studies of the large-scale structure is to account for the unseen neighbors outside the
sample window by means of the following edge-corrected estimator,
Kb DU (r) =
N X
N
V X
(r jxi
2
N i=1 j =1
fi (r)
j=
6 i
xj j) ;
(3.49)
where fi (r) is the fraction of the volume of the sphere of radius r centered on
the object i that falls within the boundaries of the sample. This estimator was
introduced by Doguwa and Upton (1989) in the field of spatial statistics.
114
2002 by CRC Press LLC
101
Kb R (r) =
N X
N
(r
V X
2
N i=1 j =1
j=
6 i
jxi xj j) ;
!
ij
(3.50)
where the weight !ij is an edge correction equal to the proportion of the area of
the sphere centered at i and passing through j that is contained in W ; in other
words, !ij is the conditional probability that the point j is observed given that it
is at a distance r from the point i. This correction is illustrated in Fig. 3.9.
Baddeley et al. (1993) give an analytic expression for !ij when W is a cube.
Note, however, that using this estimator we introduce a certain bias when we estimate n through N=V but, on the other hand, we make full use of all sample
points. For all the estimators mentioned it is possible to build the corresponding
versions for flux-limited samples by simply adding a weighting factor representing the selection function.
Stoyan and Stoyan (2000) have proposed an estimator for the K function with
a corrected (scale-dependent) intensity estimator:
N X
N
(r jxi xj j)
1 X
;
(3.51)
Kb STO (r) = e2
V (r) i=1 j =1 V (Wx \ Wy )
j=
6 i
Stein (1993) has introduced an estimator of K (r) with similar properties to
those of the Landy and Szalay estimator for (r). It is still possible that the best
dP = n3 dV1 dV2 dV3 [1 + (r12 ) + (r23 ) + (r31 ) + (r12 ; r23 ; r31 )]:
(3.52)
where (r12 ; r23 ; r31 ) is the reduced or connectedy three-point correlation function, while the quantity in square brackets is the full three-point correlation function. The terms (rij ) account for the excess of triples found as a consequence
of having more pairs than in a random distribution. Under the assumption of stationarity and isotropy is symmetric in its arguments. Peebles (2001) gives a list
y The term connected comes from their analogy with the Greens functions in particle physics.
115
2002 by CRC Press LLC
102
Figure 3.9 An illustration of the weights used in the estimator of K (Eq. 3.50) in two
dimensions. The rectangle represents the boundary of the sample. In this case, wij is the
proportion of the circumference of the circle centered at xi , passing through xj , lying
within the boundary of the sample. Depending on the relative positions of the galaxies with
respect to the boundary, different cases are illustrated: (a) wij = wji = 1; (b) wij = 1,
wji < 1; (c) wij < 1, wji < 1. It is clear from the plot that we weight the observed
neighbor xj of the galaxy xi lying at a distance r (the radius of the circle) from it by
the inverse of the probability that such a neighbor would be observed. (Reproduced, with
permission, from Martnez et al. 1998, Mon. Not. R. Astr. Soc., 298, 12121222. Blackwell
Science Ltd.)
116
2002 by CRC Press LLC
103
of the reasons that motivated people to use the three-point correlation function in
the analysis of the galaxy distribution. Eq. 3.52 is just a generalization to order 3
of Eq. 3.7. In fact, this can be done to any order n, by considering n infinitesimal
disjoint volumes dV1 , dV2 , , dVn centered at the points 1 ; 2 ; : : : ; n , then
the probability of finding a point of the point process in each of the volumes dV1 ,
dV2 , , dVn is
x x
(3.53)
where n is the product density of the n-th factorial moment measure (Stoyan,
Kendall, and Mecke 1995). If we adopt the notation (3) = for the reduced
three-point correlation function, Eq. 3.52 can be generalized to any order n, where
(4) = is the reduced cumulant four-point correlation function and (n) is the
reduced n-point correlation function. For a Poisson process (n) = 0 for n 2.
Thus far, the analysis of the galaxy catalogs by means of higher-order correlation
functions has been applied mainly to two-dimensional projected catalogs. The
counterpart of Eq. 3.52 for a catalog listing angular positions of galaxies is the
three-point angular correlation function, which is a generalization of Eq. 3.42,
dP = N 3 d
1 d
2 d
3 [1+ w(12 )+ w(23 )+ w(31 )+ z (12 ; 23 ; 31 )];
(3.54)
(3.55)
where Q is a constant. This model has been applied very successfully to galaxy
redshift catalogs. It fits the observed three-point correlation function with a value
of Q ' 1 within the range 0:1 < r < 5 h 1 Mpc.
117
2002 by CRC Press LLC
104
n (r1 ; : : : ; rn ) =
TX
(n)
t=1
Qn;t
X nY1
Ln;t
(rij );
(3.56)
where Qn;t are structure constants. In Eq. 3.56, there is a product of n 1 twopoint correlation functions. Every (rij ) corresponds to an edge rij = j i
jj
linking the n galaxies. These links are arranged in a tree structure. For every
tree, we have a product. There are T (n) distinct trees (or trees with different
topologies). For example, T (2) = 1, T (3) = 1, T (4) = 2, T (5) = 3, T (6) =
6, etc. (see Fig. 3.10). For each tree t = 1; : : : ; T (n) there are Ln;t possible
relabelings, and there is a summation of all of them. It is intriguing why this
model fits well with the observations at least up to n = 8 in the highly nonlinear
regime (strong clustering).
A simplification of Eq. 3.56 is possible if we accept that the parameters Qn;t
do not depend on the shape of the tree,
x x
n (r1 ; : : : ; rn ) = Qn
TX
(n)
t=1;Ln;t
(rij ):
(3.57)
A more general model that verifies Eq. 3.56 is the scale-invariant model proposed by Balian and Schaeffer (1989),
(3.58)
118
2002 by CRC Press LLC
105
Figure 3.10 The different shapes of trees connecting n points. The number of possible relabelings is indicated in parentheses. (Reproduced, with permission, from Coles and Luchin
1995, Cosmology. The Origin and Evolution of Cosmic Structure. c John Wiley & Sons).
Z
A
dx1 : : :
Z
A
dxn n (x1 ; x2 ; : : : ; xn );
(3.61)
i.e., the expression above is the factorial moment of order n of the random variable
(A). Eq. 3.61 is read in spatial statistics by stating that the density of the factorial
moment n with respect to the Lebesgue measure is the product density of order
n n . In fact, Eq. 3.61 encapsulates the relation between the n-point correlation
functions and the count in cell probabilities (White 1979).
As high-order correlation functions depend on a large number of arguments,
they are rather difficult to estimate. Counts in cells are easy to estimate; Szapudi
(1998) has developed an ingenious exact algorithm for the computation of counts
in cells. The expected count distributions can be predicted theoretically, thus they
are frequently used to study higher correlations in the galaxy distribution.
119
2002 by CRC Press LLC
106
For that we have to establish the connection between the counts and correlation
functions (see, e.g., Szapudi and Szalay 1993; Szapudi, Colombi, and Bernardeau
1999). This is similar to the relation (3.61) above. The variance of the counts is
the average of the correlation function in a cell:
1
= 2
V (A)
Z
A
A
(x1 ; x2 ) d3 x1 d3 x2 r:
Using the hierarchical assumption (3.57), the first structure constants are
Q2 = 1, and the higher-order constants (connected moments) are
1
1
Qn = n 2 n 1 n
n V (A)
A
:::
A
Q1 =
n (x1 ; : : : ; xn ) d3 x1 : : : d3 xn :
Sn = nn 2 Qn = nn 1 :
The structure constants Q can be calculated using the factorial moments defined
above. The first four connected moments are:
n = 1 ;
= 22 1;
1
( 31 2 + 213 )
Q3 = 1 3
;
3(2 12 )2
2 ( 43 1 322 + 122 12
Q4 = 1 4
16(2 12 )3
614 )
The skewness and kurtosis are usually described by the parameters S3 and S4 .
The sample errors of these parameters are thoroughly analyzed in Szapudi, Colombi, and Bernardeau (1999); a program package (FORCE) for calculating the
errors can be found at I. Szapudis Web page (Szapudi). As the parameters Q are
given by models of the correlation hierarchy, all these models can be tested using
counts in cells.
The formalism has been mainly applied to photometric catalogs (projected
data). Gaztanaga (1994) estimated the parameters S and Q up to the ninth order,
using the APM galaxy catalog. Because there are far fewer galaxies in redshift
catalogs, the only moments estimated using three-dimensional counts are skewness and kurtosis; the sample errors for higher moments are too large. Fig. 3.11
shows an example of the skewness and kurtosis of the PSCz redshift survey, found
by Szapudi et al. (2000). As we see, the skewness and kurtosis practically do not
depend on scale (the depth of a subsample), and their values are S3 = 1:9 0:6
and S4 = 7:0 4:1, in agreement with previous studies.
120
2002 by CRC Press LLC
107
Figure 3.11 Skewness (upper panel) and kurtosis (lower panel) for the PSCz redshift survey. Different points describe different volume-limited subsamples. The lines show the predictions of a galaxy formation model proposed by Benson et al. 2000; solid lines show the
mean and dotted lines show the variance. (Reproduced, with permission, from Szapudi et
al. 2000, Mon. Not. R. Astr. Soc. 318, L45L50. Blackwell Science Ltd.)
F (r) = P (X r) = 1 P (0; Br )
121
2002 by CRC Press LLC
108
F (r) (also referred as to Hs (r) in spatial statistics) is the spherical contact distribution function. For a homogeneous Poisson process of intensity , it is given
by
4 3
r :
3
N (r)
Pb(0; Br ) = emp
Nb
This is basically equivalent to the minus-sampling estimator (Baddeley et al.
1993) of the spherical contact distribution function, which reads
Fb (r) =
V fp 2 A : S (p; r) A ^ mini jp
V fp 2 A : S (p; r) Ag
xij rg ;
(3.62)
where f i gN
i=1 are the observed points (galaxies) and S ( ; r ) is the sphere centered at random test point and having radius r. This estimator is unbiased.
Fig. 3.12 shows the function F (r) estimated for a volume-limited sample extracted from the CfA1 catalog with 80 h 1 Mpc depth, together with the same
function for a binomial process having the same number of points.
If we arbitrarily place a point within the boundaries of a realization of a clustered point process, it is likely that it lies farther away from an event than it would
in the case of placing it over on a homogeneous Poisson process. Therefore, for a
clustered pattern at scale r, we should expect to obtain F (r) < FPois (r).
Note that Eq. 3.62 provides an estimator of 1 P0 (x), i.e., 1 minus the void
probability function. White (1979) has shown how the void probability function
serves as a generating functional for all the volume probabilities P (N; V (A)),
and hence it is related to the n-order correlation functions.
109
Figure 3.12 The F -function for a galaxy sample (dashed line). The shaded area corresponds to the rms fluctuations of the same function calculated over 15 realizations of a
binomial process with the same number of points. (Reproduced, with permission, from
Kerscher and Martnez 1998, Bull. Int. Statist. Inst., 57-2, 363366.)
function G(r) of the distance between events of the process (galaxies in our case),
i.e., G(r) is the probability that the distance between a randomly chosen galaxy
and its nearest neighbor (event) is less than or equal to r.
An indication of clustering will be that G(r) > GPois (r), because the distance
from clustered points to their nearest neighbor usually will be shorter than in a homogeneous Poisson distribution. Likewise, for a regular pattern G(r) < GPois (r),
because a point from such a process is, on average, farther away from its nearest
neighbor than in a Poisson process. This is illustrated in Fig. 3.13.
It is clear from the definitions that for a homogeneous Poisson process G(r) =
F (r). In general, however, G(r) does not necessarily verify any continuity property (Baddeley 1999). The main drawback of nearest neighbor statistics is that
they provide information of the clustering pattern within a rather restricted range
of scales (Cressie 1991). Nevertheless, within these scales they are extremely useful. In fact, the quotient
J (r ) =
1 G(r)
;
1 F (r)
(3.63)
suggested by van Lieshout and Baddeley (1996), has provided rather interesting
results when applied to galaxy data and mock galaxy catalogs drawn from N body simulations (Kersher et al. 1999). By means of this quotient we can compare
the region around an event (galaxy) of the point process with the events lying
within the neighborhood of a randomly chosen point. For a homogeneous Poisson
123
2002 by CRC Press LLC
110
Figure 3.13 The nearest neighbor distribution function and the J -function for a galaxy
sample (dashed line). The shaded area corresponds to the rms fluctuations of the same
functions calculated over 15 realizations of a binomial process with the same number of
points. (Reproduced, with permission, from Kerscher and Martnez 1998, Bull. Int. Statist.
Inst., 57-2, 363366.)
111
Measurements of the luminosity or morphological segregation have been attempted by applying statistical descriptors, such as the two-point correlation function, the nearest neighbor distribution functions, or multifractal measures to subsamples of the survey corresponding to populations of a given type, for example,
all elliptical galaxies, or all galaxies brighter than a particular luminosity cutoff
(Hamilton 1988; Davis et al. 1988; Salzer, Hanson, and Gavazzi 1990; Domnguez-Tenreiro, Gomez-Flechoso, and Martnez 1994; Loveday et al. 1995; Hermit
et al. 1996; Guzzo et al. 1997). When the different populations provide different
results for the statistical descriptors, showing different clustering properties at
given scales, we can interpret this result as the fingerprint of segregation.
Other studies have been based on the cross-correlation function of subsamples
(Alimi, Valls-Gabaud, and Blanchard 1988). The cross-correlation function between two types of objects, namely A and B, is defined similarly to Eq. 3.7 by
considering the joint probability of finding an object of type A in the volume element dVA , and an object of type B in the volume element dVB , both volumes
separated by a distance r:
Beisbart and Kerscher (2000) have demonstrated the use of some statistical
measures to quantify the mark correlation properties of point processes in the
analysis of the galaxy distribution. Here we describe one of them: the normalized
mark correlation function.
125
2002 by CRC Press LLC
112
Figure 3.14 The normalized mark correlation function for a volume-limited subsample
of the SSRS2 catalog. The shaded area shows the rms fluctuations calculated over 1000
realizations of randomly shuffling the marks. (Reproduced, with permission, Beisbart and
Kerscher 2000, Astrophys. J., 545, 625. AAS.)
m
=
m(m)dm:
126
113
x x
M
2 ((x1 ; m1 ); (x2 ; m2 ))dV1 dm1 dV2 dm2
as the joint probability that in the volume element dV1 lies a galaxy with the mark
in the range [m1 ; m1 + dm1 ] and another galaxy lies in dV2 with the mark in
[m2 ; m2 + dm2 ]. Then the Stoyan mark correlation function for an isotropic and
1
kmm (r) = 2
m1 m2 M
2 ((x1 ; m1 ); (x2 ; m2 ))dm1 dm2 ; (3.64)
m
2 (r)
for 2 (r) 6= 0.
Intuitively kmm (r) can be considered the squared geometric mean of the marks
normalized by the overall mean mark squared, under the condition that the pairs
of galaxies are separated by r. kmm (r) < 1 represents inhibition of the marks at
the scale r. For example, in forests it is typically found that trees with larger stem
diameter (mark) tend to be isolated. Using luminosity as the mark, the opposite
effect has been found for the galaxy distribution, i.e., kmm (r) > 1 at small scales
(Beisbart and Kerscher 2000), implying stronger clustering of brighter galaxies at
small separations (see Fig. 3.14).
Other related measures such as the mark covariance function (Stoyan 1984b;
Cressie 1991) and the mark variogram (Walder and Stoyan 1996) can be used
to analyze a marked point process. More details and practical formulae for the
estimators for both continuous and discrete marks can be found in Stoyan and
Stoyan (1994), x 15.4.4, and Beisbart and Kerscher (2000).
127
2002 by CRC Press LLC
CHAPTER 4
116
The strongest observational evidence supporting the validity of the cosmological principle is the isotropy of the cosmic microwave background radiation. Other
observations supporting this hypothesis include the angular distribution of radio
sources, the analysis of the X-ray background, and the distribution of quasars and
Lyman- clouds. The distribution of
-ray bursts also supports the homogeneity picture. These are all essentially two-dimensional tests because they analyze
objects or radiation as seen on the celestial sphere. Additional observational evidence of this kind comes from the analysis of the angular correlation function
for the Lick and the APM galaxy surveys. This function scales with magnitude as
predicted for a homogeneous universe (Peebles 1993). In this chapter we analyze
these tests.
The fractal or scaling approach is mathematically rigorous and applicable to
measuring the galaxy clustering (Martnez 1991). In this chapter we explain how
to measure fractal dimensions for the galaxy catalogs, how these quantities are
related to other statistical descriptors, and the physical implications of the fractal prescription. Given that cosmic clustering has evolved under the influence of
gravitation only, there is no physical motivation to think about preferred scales in
the galaxy formation process. In this context multifractal measures are described
as the most natural scaling behavior of the moments of the counts in cells introduced in Section 3.6. Recent views related to fractal aspects of galaxy clustering
as the multiscaling prescription are discussed in this chapter.
4.2 Fractal models for the universe
Several fractal constructions have been proposed thus far as artificial universes
trying to mimic some properties of the real galaxy distribution. Mandelbrot (1975)
proposed a model based on a RayleighLevy flight, where galaxies are placed at
the steps of a random walk. The direction of each jump is isotropically chosen at
random and its length follows a power-law probability distribution function.
4.2.1 RayleighLevy dust
This model, based on the RayleighLevy flight, is built as follows. Choose a starting point of the simulation. Following a random walk, jump over to a second
point, and so on. The power-law constraint affects only the distribution of the
jump lengths, while the direction of each jump is taken isotropically at random.
If X denotes the random variable that contains the values of the jump lengths, its
probability distribution function must be a power law
P (X > ) =
0
D
(4.1)
If we place a galaxy at each jump the output appears as shown in Fig. 4.1. The density of this distribution of points within a sphere of radius R varies as RD 3 . This
129
2002 by CRC Press LLC
117
Figure 4.1 Simulation of a RayleighLevy flight in two dimensions. Several zooms are
shown to illustrate the self-similarity of this construction. (Reproduced, with permission,
from Nusser and Lahav 2000, Mon. Not. R. Astr. Soc., 313, L39L42. Blackwell Science
Ltd.)
118
Figure 4.2 This diagram illustrates how the Soneira and Peebles fractal model is built in
two dimensions. The left panel shows the construction when overlapping of the spheres
in each level is avoided. The right panel shows the opposite case when overlapping is
allowed.
will be studied in detail, both qualitative and quantitatively, in the last section of
this chapter.
4.2.2 Soneira and Peebles fractal model
Soneira and Peebles (1978) introduced a model based on the superposition of
fractal clumps, which reproduced not only the appearance of the projected distribution of galaxies as seen in the Lick maps (Seldner et al. 1977), but also the
angular two- and three-point correlation function. Each fractal clump is built as
follows: in a sphere of radius R, we randomly place spheres of radius r=,
with > 1. Within each of these spheres, we again place new spheres of radius R=2 . The process is repeated L times and the last generation of L centers
are considered the galaxies of a clustering hierarchy or a bounded fractal with
dimension D = log = log .
Fig. 4.2 shows several steps of the construction of this fractal pattern in two
dimensions.
What does this fractal clump look like when projected onto the sky? Peebles
(1998) showed a series of projections of the particles lying in concentric shells as
seen by an observer situated at a given point close to the center of the first sphere.
A similar set of equal-area HammerAitoff projections is shown in Fig. 4.3. This
is a fractal clump with D = 2, ( = 2, ' 1:41, and L = 18). We have scaled
the fraction of particles plotted as a function of 1=r2 , with r as the distance of a
particle from the center. The width of a given shell is always twice the width of
the previous one. The anisotropy of the point distribution provided by this model
is remarkable. The inhomogeneity remains even in the larger shells far away from
131
2002 by CRC Press LLC
119
Figure 4.3 Left panel: Equal-area HammerAitoff projection of a single Soneira and Peebles fractal clump as seen by an observer situated at a point of the model close to the
center of the first sphere. In each panel, from top to bottom, we have projected the points
lying in concentric shells of increasing radius and width: 1:25 < r
2:5, 2:5 < r
5,
5 < r
10, 10 < r
20, in arbitrary units. Right panel: The same projections but for
the IRAS 1.2Jy redshift survey. The division in shells is the same as before, but now the
units are thousands of km/s in the radial velocity of each galaxy.
132
2002 by CRC Press LLC
120
the center. Davis (1997) performed the same kind of projection for a real fluxlimited redshift survey, the IRAS 1.2Jy catalog. The right panels of Fig. 4.3 show
the IRAS galaxies lying in concentric shells of increasing size projected onto the
sky. The radial distribution of shells is the same as that shown for the Soneira
and Peebles model (left panels in Fig. 4.3), but it is slightly different from that
shown in Davis (1997). We can see that for the more distant and larger shells
the distribution appears more homogeneous. This tendency toward homogeneity
detected visually in the redshift catalogs will be quantified in the next sections.
4.3 Tests on projected data
The nearby distribution of galaxies is fairly inhomogeneous and can be considered
fractal, but farther away the distribution appears to become more homogeneous.
As described in Chapter 2, direct tests for homogeneity based on galaxy redshift
catalogs do not yet provide definite answers. The problem lies in the unknown
luminosity evolution of galaxies that can change selection limits and affect catalog
densities at large distances.
A stronger result can be obtained by galaxy number counts (see also Chapter 2).
Their dependence on the magnitude limits is also affected by galaxy luminosity
and number density evolution, but only at larger distances. At closer distances
the number counts behave exactly as they should for a uniform D = 3 density
distribution (see the review in Peebles 1993).
Further evidence for a nonfractal D = 3 distribution of galaxies is provided by
the scaling of the angular correlation function of galaxies w() (see Chapter 8).
The scaling predicted by a nonfractal distribution,
w() =
1
W (D);
D
121
Figure 4.4 Distribution of radio sources. (Reproduced, with permission, from Baleisis et
al. 1998, Mon. Not. R. Astr. Soc., 297, 545558. Blackwell Science Ltd.)
Figure 4.5 Distribution of
-ray bursts (courtesy of NASA Marshalls Space Flight Center,
Space Sciences Laboratory).
role. Fig. 4.5 shows the distribution of directions to gamma-ray bursts. This is also
isotropic, but as yet nothing is known about the distances to their sources.
The maps listed above are isotropic. Although it is difficult to generate isotropic
fractal maps (typical maps are shown in the previous section), it could be possible.
Durrer et al. (1997) present fractal constructions that are more or less isotropic in
projection. These are, however, rather special fractals and it is not clear if similar
constructions could be used for random fractals.
The spatial distribution of optical and radio galaxies, and X-ray and gammaray sources could also be modified by spatially fluctuating biasing. The most
isotropic background of all, the cosmic microwave background (CMB), measures, according to the present cosmological theory, fluctuations of total matter
density at the moment of recombination. The typical CMB temperature fluctuations T=T 10 5 translate into fluctuations of the same amplitude for =
134
2002 by CRC Press LLC
122
Figure 4.6 Density fluctuations as a function of the scale for different models and different
observations, including galaxy surveys, radio sources, X-ray background, and cosmic microwave background data. (Reproduced, with permission, from Wu, Lahav, and Rees 1999,
Nature, 397, 225235. Macmillan Publishers Ltd.)
at scales of the horizon size. This value is impossible to reconcile with the fractal
picture.
Present knowledge about density fluctuations on different scales is summarized
in Fig. 4.6. The rms density fluctuation amplitudes are small already on the scales
of a few tens of Mpc and diminish rapidly toward larger scales. Thus the fractal
picture is probably useful only for smaller scales. The density distribution at large
scales is fairly homogeneous. Although a totally fractal universe would be an
enormous challenge both for observers to map out and for theoreticians to explain,
the more modest task of discovering how the locally fractal distribution changes
into a homogeneous one is also very interesting.
4.4 Fractal dimensions
4.4.1 Hausdorff dimension
Let us consider a nonempty subset A of the Euclidean space IRd . For r > 0, an
r-cover of A is a countable collection of sets with diameter less than or equal to
135
2002 by CRC Press LLC
Fractal dimensions
123
H (A) = 1
and
H (A) = 0
< DH (A)
(4.4)
> DH (A):
(4.5)
if
if
Therefore, the Hausdorff measure has a critical behavior at the Hausdorff dimension, jumping from infinity to zero.
Mandelbrot defined a fractal set as one for which the Hausdorff dimension
strictly exceeds its topological dimension.
It is rather straightforward to show that the Hausdorff dimension of any countable set is zero (Falconer 1990); therefore, this measure really is not useful for
application to the galaxy distribution. It has been attempted, however, on the basis of the following assumption: If galaxies were good tracers of mass and the
mass distribution had a Hausdorff dimension less than 3, could we estimate this
value just from the analysis of the galaxy distribution?
Let us put this question in a different way using an example from the field of
complex systems. We know that the Lorenz attractor is a fractal structure in 3-D
with the Hausdorff dimension DH ' 2:1. The previous question, in this context,
should read as follows: If we know the coordinates of N points distributed randomly on the three-dimensional trajectory of the attractor, should we be able to
calculate the attractor Hausdorff dimension just from the point distribution? This
has been done by Domnguez-Tenreiro, Roy, and Martnez (1992) by means of
the minimal spanning tree (MST). This is a graph-theoretical construction linking all points of the process without closed loops and with minimal total length
(van de Weygaert, Jones, and Martnez 1992). In Chapter 10, we provide a more
formal definition and see how this construction has been used as a measure of
the structure of the galaxy distribution. Here, the MST plays the role of the minimal covering, implicitly given by the infimum requirement in the definition of
the Hausdorff dimension (4.3). The method works as follows: Given N points
sampling the attractor we randomly choose NR points and calculate the lengths
of the m = NR 1 branches of the MST connecting all those points f`i gm
i=1 (see
136
2002 by CRC Press LLC
124
Figure 4.7 The left panel shows 1000 points sampling the trajectory on the Lorenz attractor; the right panel shows its MST. (Reproduced, with permission, from DomnguezTenreiro, Roy, and Martnez 1992, Prog. Theor. Phys., 87, 11071118.)
S (m) =
m
1X
` :
m i=1 i
S (m) = K ( )m =h() ;
the fixed point of the function h( ) provides a good estimator of the Hausdorff
dimension of the attractor. Using this method Domnguez-Tenreiro, Roy, and
Martnez (1992) estimated the Lorenz attractor Hausdorff dimension to be DH =
2:155 0:005. A similar algorithm was applied by Martnez and Jones (1990) for
the galaxy distribution. The value obtained by them for the CfA1 catalog was 2.1.
Fractal dimensions
125
defined as
DC = lim
!0
log N ()
:
log(1=)
(4.6)
DC '
d log N ()
d log(1=)
(4.7)
3
;
4
(4.8)
where Ncel () = V (A)=3 is the number of cuboidal cells of side that can be
put inside A.
In this way we are provided with a new procedure to estimate D0 by replacing
the above expression for N ()v in (4.6) or in (4.7). For a homogeneous Poisson
process with intensity , the situation is the following:
V (A)
1 exp( 3 ) :
(4.9)
3
Obviously, for a Poisson process, at large scales N () / l 3 but, if is small,
N () =
the exponential term will dominate at small scales. So, although the method introduced is correct, it can only be applied to samples with high enough number
density and at long enough distances.
4.4.3 Correlation dimension
One of the most useful statistical measures to study fractality on large scales is
the integral of the correlation function:
N (< r) = n
Z r
4s2 (1 + (s))ds:
138
(4.10)
126
This function measures the number of neighbors on average that a given galaxy
has within a distance r. A point pattern is said to be fractal with the correlation
dimension D2 when
N (< r ) / r D2 :
(4.11)
The correlation dimension is evaluated as:
D2 =
(4.12)
1 r 0
g(r )dV;
V 0
and the structure function g (r) = 1 + (r). As we see, gb(r) is the volume mean
of Ripleys K -function, introduced in Chapter 3.
The two-point correlation function for a fractal pattern should vary as 1 +
(r) / r3 D2 . However, there are problems when we try to apply standard meas-
gb(r) =
( r) =
hn(x)n(x + r)i ;
n
(4.13)
(r) =
(r)
n
1;
depends, if measured for a fractal, on the sample size (for a fractal of dimension
D the mean density of a sample n RD 3 , where R is the depth of the sample).
If we try to verify the hypothesis of fractality of the galaxy distribution, we cannot use directly the usual border correction methods, as these assume the homogeneity of the parent population. Thus, the estimators for the conditional density
(r) have to be chosen among minus-estimators, which use only those galaxies
around a galaxy that are closer to that specific galaxy than the borders of the sample. This eliminates most of the sample points as centers for larger distances r and
results in large sample variance. In any case, if edge-corrected estimators must be
applied, caution has to be exercised to avoid the introduction of spurious homogeneity. We recommend applying the estimators to specific fractal point patterns
to test this possibility. We return to this point in Section 4.4.5.
139
2002 by CRC Press LLC
Fractal dimensions
127
3 r
(s)s2 ds
= 3
r 0
(4.14)
is less noisy, at the expense of smoothing the possible details in (r). For a simple
fractal, however, it is also a power law, ? (r) R3 D .
Pietronero and his colleagues applied these relations to galaxy catalogs in various papers (for a review see Sylos Labini, Montuori, and Pietronero 1988). They
have always found a perfect power law for (r) and ? (r) for the full argument range, confirming the fractal nature of the galaxy distribution. The left panel
of Fig. 4.8 shows a typical result for the StromloAPM galaxy redshift survey.
This can be compared with a later careful re-analysis of the same survey, using
the same methods (Hatton 1999; the right panel of Fig. 4.8). The depths of the
volume-limited samples, selected from the survey, are somewhat different, but
the main difference is in the scale range chosen to determine the fractal dimension. Hatton showed that the results at small scales are severely contaminated by
the Poisson noise. Eliminating these scales from consideration gives the fractal
dimension that approaches the uniform D = 3, as the depths of the samples grow.
Crossover to homogeneity also has been found in other studies (Martnez, in
press). In Fig. 4.9, we show the function 1 + (r), estimated with proper care, for
several deep galaxy redshift surveys (Guzzo et al. 1991; Bonometto et al. 1993;
Martnez 1999). The fractal behavior at small scales disappears at larger distances,
providing evidence for a gradual transition to homogeneity (Einasto and Gramann
1993). It is remarkable that the break of the scale-invariant power law appears at
the same scale, approximately 10 to 15 h 1 Mpc, for the four samples. Martnez
et al. (1998) have studied the correlation dimension by means of Eq. 4.11 for the
StromloAPM survey, and have found a scale-dependent behavior of D2 . This
quantity reaches a value of 2.8 at the largest scales probed by the sample in agreement with the results by Hatton (1999).
A scale-dependent correlation dimension was also found by Amendola and
Palladino (1999) estimating the function gb(r) from the Las Campanas survey.
These authors have developed an interesting method to reliably measure gb(r) at
large distances based on radial cells, reporting values of D2 around the homogeneity value 3 for the largest analyzed scales. In a recent paper, Pan and Coles
(2000) studied the PSCz catalog by means of the multifractal algorithms. They
obtained a value of D2 = 2:99 at scales larger than 30 h 1 Mpc, while for r
below 10 h 1 Mpc, they obtained D2 = 2:16, a value comparable with that
obtained by Martnez and Coles (1994) for the QDOT-IRAS galaxy redshift survey. The transition to homogeneity is also apparent in the plots of ? (r) obtained
by the proponents of the unbounded fractal universe. In Sylos Labini, Montuori,
and Pietronero (1998), the plots of ? (r) corresponding to samples such as the
StromloAPM or the IRAS 2Jy show an unambiguous deviation from a power
140
2002 by CRC Press LLC
128
vl18
vl19
vl20
D=2.04
D=2.11
D=2.1
10
(r)
10
10
10
10
1
r (h Mpc)
100
Figure 4.8 The conditional average densities for volume-limited samples of different depth
from the StromloAPM redshift survey. Observational results are denoted by dots and the
respective power laws by lines (see the legends). Lower lines correspond to deeper samples, and the fractal dimensions for different samples are denoted by D . Errors are estimated by a bootstrap procedure. The left panel shows the results of Sylos Labini and
Montuori (1998), and the right panel shows the results of Hatton (1999). (Reproduced,
with permission, from Sylos Labini and Montuori 1998, Astr. Astrophys., 331, 809814
and from Hatton (1999), Mon. Not. R. Astr. Soc., 310, 11281136. Blackwell Science Ltd.)
law at large scales. Despite the fact that the authors interpreted these results in a
different way, they could likely indicate the transition to homogeneity.
Fractal dimensions
129
Stromlo-APM
Las Campanas
ESP
10
CfA2
1+(r)
D2= 2
D2= 3
10
100
-1
r (h Mpc)
Figure 4.9 The function 1 + (r ) for several redshift surveys (see the legend). Two reference lines have been plotted, one corresponding to a fractal with dimension D2 = 2
and the other corresponding to a homogeneous distribution (D2 = 3). We can appreciate
how all data show a gradual transition from the fractal regime at short scales to a more
homogeneous distribution at large scales.
D r
(r ) =
3 Rs
D
1;
D 1=(3 D)
Rs
6
is proportional to the depth of the sample Rs .
r0 =
This prediction has been tested on simple fractals, such as the one introduced in
the previous section (Paredes, Jones, and Martnez 1995; Martnez, Lopez-Mart,
and Pons-Bordera 2001). Fig. 4.10 shows the behavior of r0 with sample radius
for spherical sample of a fractal clump generated by means of the Soneira and
Peebles model introduced in Section 4.2.2.
We show here how this prediction is not supported by the observed galaxy
distribution (Martnez et al. 1993; see Martnez, Lopez-Mart, and Pons-Bordera
2001 for details). The analyses have been performed on the CfA2 redshift catalog
(Geller and Huchra 1989; Park et al. 1994). From the CfA2 north survey, we have
142
2002 by CRC Press LLC
130
14
-1
r0(Rs) (h Mpc)
12
10
8
6
4
2
0
0
10
15
20
-1
Rs (h Mpc)
25
30
35
Figure 4.10 The correlation length as a function of the sample radius for a fractal
point distribution. (Reproduced, with permission, from Martnez, Lopez-Mart, and PonsBordera 2001, Astrophys. J. Lett., 554, L5L8. AAS.)
Fractal dimensions
131
4:21 0:31
6:65 0:29
6:43 0:30
6:88 0:32
6:88 0:49
Figure 4.11 Volume-limited samples extracted from the CfA2 north survey. From left to
right and top to bottom the depth of the samples increases. The size of the diagrams varies
according to the real volume of the sample. The number of points in each one is, in the
same order (left to right, above), 736, 1113, 1159, 1134, and 905. The correlation length
r0 of each sample is shown together with the standard deviation in the weighted loglog
linear regression.
132
r0
3
35
40
45
50
55
60
Rs (h-1 Mpc)
65
70
75
Figure 4.12 The correlation length as a function of the sample size for several volumelimited subsamples drawn from the CfA2 redshift catalog. (Reproduced, with permission,
from Martnez, Lopez-Mart, and Pons-Bordera 2001, Astrophys. J. Lett., 554, L5L8.
AAS.)
Fractal dimensions
133
Figure 4.13 (a) A realization of the diffusion-limited aggregation process. (b) The density
within circles of radius R as a function of R. (c) The function 1 + (r ) calculated for
the subsamples with different outer radius R. (Reproduced, with permission, from Paredes,
Jones, and Martnez 1995, Mon. Not. R. Astr. Soc., 276, 11161130. Blackwell Science
Ltd.)
134
for a set in IRd . Although this relation is less accurate than the one used in the
definition of the correlation dimension, in which we average considering all the
galaxies in the sample as possible centers, it has the advantage that it permits us
to extend the measure of the dimension to much larger scales, due to the geometry
of redshift surveys typically centered at the observer. We can illustrate the use of
this dimension with a typical aggregation model studied in growth phenomena. It
is the well-known diffusion-limited aggregation (DLA) introduced by Witten and
Sander (1981). This process is simulated by setting a seed particle at the origin of
a lattice. A new point moves around randomly until it visits a site in contact with
the seed. Then its location is fixed and the particle becomes part of the cluster.
Subsequent particles are introduced in the same way until they join the cluster.
The fractal cluster shown in Fig. 4.13 is the output of one DLA realization. It
is obvious from the picture that the density in concentric circles of radius R is a
decreasing function. In the figure we show how the density varies, falling off as a
power law n(R) / RDM 2 where DM is the massradius dimension. Performing
the calculation of (r) in several discs with different radius, we can see how the
variation of density affects the amplitude of the correlation function. Fig. 4.13
shows the power-law shape of the structure function (1 + (r)) / r
for the
DLA subsamples marked with increasing radius. We can see how the smaller the
density, the larger the amplitude of the structure function, while the slope always
remains the same,
= 2 DM . As explained in Section 4.4.4, this effect is a
consequence of the density decrease with volume of the fractal pattern.
4.5 Multifractal measures
Multifractals have been applied to the description of chaotic dynamical systems in
recent years. Multifractals are fractal sets with an invariant measure characterized
by a whole spectrum of singularities, instead of a single number as in homogeneous fractals. The theory of multifractals was developed based on the pioneering
work of Renyi in the framework of the probability theory (Renyi 1970). In this
context the Renyi dimensions have been used in the description of strange sets
by Hentschel and Procaccia (1983). However, a different approach came from a
generalization of the classical concept of Hausdorff measures. This way of introducing the multifractal measures in the context of the analysis of nonlinear
dynamical systems has been employed by Halsey et al. (1986), who noticed that
the Hausdorff generalized dimensions are not always equivalent to the Renyi dimensions.
To illustrate a multifractal measure, we introduce here the model described
by Meakin (1987). The construction of this set is performed following a multiplicative cascade (see also Martnez et al. 1990). We start dividing a square
147
2002 by CRC Press LLC
Multifractal measures
135
into four square pieces. We assign a probability number ffi g4i=1 to each piece
P4
( i=1 fi = 1). Each of the small squares is again subdivided into four pieces, assigning to each of the subsquares one of the numbers fi randomly permuted. The
probability attached to each one of the new 16 squares is the product of this number and the number of its parent. We continue this construction so on and so forth,
each time multiplying the corresponding number fi from the random permutation
by all its ancestors. If we perform the previous construction until L = 8 levels,
we have at the end a 2562 lattice with a measure associated with each pixel. As
an illustration we have performed several realizations of this process for different
choices of the initial parameters:
Model
f1
f2
f3
f4
I
II
III
IV
1/3
0.448
0.5714
0.22
1/3
0.3
0.2857
0.24
1/3
0.2
0.1429
0.26
0
0.052
0
0.28
Z (q; ) =
N
()
X
i=1
(Bi ())q
(4.15)
where (Bi ()) is the measure associated with the box of size , Bi (). The Renyi
dimensions are defined by
Dq = lim
!0
log Z (q; )
(q 1) log
(4.16)
Dq = (1 q) 1 log2 @
4
X
i=1;fi 6=0
fiq A :
136
Figure 4.14 Multifractal multiplicative cascades with parameters given in the text. Models
are I, II, III, and IV, from top to bottom and left to right.
with D0 ' 1:58 and D2 ' 1:22 and Model IV is, like Model II, a multifractal
measure on a nonfractal support, the unit square. Fig. 4.15 shows the curves Dq
for the four models. For the simple fractal, Dq is a straight line with zero slope.
Multifractals present characteristic nontrivial shapes for Dq .
The previous approach is a generalization of the box-counting algorithm to
define the function Dq . Likewise, it is often more efficient to generalize the correlation algorithm to obtain an expression for Dq . Formally the generalized dimensions defined in the following paragraphs are not completely equivalent, but we
use the same notation as before for the dimensions, with the scale now represented
by r. The scaling property used to define the correlation dimension hN ir / rD2
might be generalized to moments of any order if
Z (q; r) =
N
1X
n (r)q 1 / r (q) ;
N i=1 i
(4.17)
Multifractal measures
137
5
I
II
III
IV
4.5
4
3.5
Dq
3
2.5
2
1.5
1
0.5
0
-20
-15
-10
-5
10
15
20
Figure 4.15 The Renyi dimension for the multiplicative multifractal measures corresponding to Models I, II, III, and IV.
W (; n) =
N
1X
r (n)
N i=1 i
/ n1 q :
(4.18)
For q = 0, the previous equation provides a way for estimating the capacity
dimension D0 . Galaxy samples such as the CfA1 provide a value of D0 ' 2:1,
while the correlation dimension of the same sample is D2 ' 1:3 in the range of
scales 1 r 10 h 1 Mpc (Martnez et al. 1990; Jones, Coles, and Martnez
150
2002 by CRC Press LLC
138
Figure 4.16 The Renyi dimensions for the Abell and ACO clusters. (Reproduced, with permission, from Borgani et al. 1994, Astrophys. J., 435, 3748. AAS.)
1992). Eq. 4.17 is best suited for q 2, while Eq. 4.18 works better for q < 2.
Clusters of galaxies also present scaling behavior, although the Dq values are
different from the Dq values for galaxies. Borgani et al. (1994) have shown that
the correlation dimension turns out to be D2 ' 2 for two different samples of
clusters of galaxies within the range 15 r 60 h 1 Mpc. The difference in the
values of the correlation dimension for galaxies and for clusters of galaxies may
be interpreted under the multiscaling hypothesis (Jensen, Paladin, and Vulpiani
1991), as shown in next section. The entire Dq function for clusters of galaxies is
shown in Fig 4.16.
There is a clear connection between the multifractal approach and the moments
of the counts in cells (Borgani 1993). In fact, the partition sum defined from the
moments of the box-counting measures in (4.15) can be easily written in terms of
the probabilities P (N; V ) introduced in Section 3.6:
N
tot
X
k q
P (N; 3 );
N
tot
k=1
where Ntot is the total number of points and Ncel is the total number of cells. This
approach has been exploited by Borgani (1993, 1995) who has obtained the Dq
Z (q; ) = Ncel ()
function for several point distributions and stochastic models with known expressions for the probabilities P (N; V ).
4.6 Multiscaling
Integral quantities such as
large scales. For r 10 h
Multiscaling
139
Figure 4.17 The multifractal cascade model in 3-D (a), and after applying two different
density thresholds (b and c). (Reproduced, with permission, from Martnez 1996, in Dark
Matter in the Universe, IOS Press, Societ`a Italiana di Fisica.)
the distribution of clusters of galaxies. In this section, we will provide an explanation of the clustering of different objects such as galaxies or clusters within the
same theoretical framework.
If the matter distribution is considered a continuous density field, we could
think of galaxies as the peaks of the field above some given threshold. A larger
threshold will correspond to clusters of galaxies. The higher the threshold, the
richer the galaxy cluster. We can use the multiplicative model introduced in the
previous section to illustrate this behavior. We build the model in 3-D by dividing
a cube of side 100 h 1 Mpc into eight equal cubic pieces of side 50 h 1 Mpc.
A probability number fi is assigned to each part. The subdivision process is repeated L times. Jones, Coles, and Martnez (1992) have chosen f1 = f2 = 0:07,
f3 = 0:32, f4 = 0:54, and fi = 0 for i = 5; : : : ; 8. At the final stage we end
with a lattice of 2L 2L 2L cells, each with an assigned measure (density)
fi1 fi2 fiL resulting from the multiplicative cascade. The parameters of the
model were chosen to provide values for D0 = 2:0 and for D2 = 1:3. Point
distributions are generated on the cells of the lattice according to their densities.
Small random shifts are applied to the position of the points to avoid regularity
effects resulting from the lattice. By applying different density thresholds, which
are quite naturally defined in this model, we obtain the distributions shown in
Fig. 4.17.
The value of D2 is approximately the slope of the loglog plot Z (2; r) versus
r, shown as a solid line in Fig. 4.18. After applying the threshold, Z (2; r) still
follows a power law, but with different slope (see the dotted and dashed lines
in the figure). In the plot we can see that the higher the density threshold, the
lower the value of D2 . Multiscaling is a scaling law where the exponent slowly
varies with the length scale due to the presence of a threshold density defining the
objects.
152
2002 by CRC Press LLC
140
Figure 4.18 The function Z (2; r ) for the multiplicative model. Note that Z (2; r ) is normalized here differently than in Eq. 4.17. Here ni (r ) has been replaced with ni (r )=Ntot .
Lower slopes are obtained when the density threshold is increased. (Reproduced, with permission, from Martnez 1996, in Dark Matter in the Universe, IOS Press, Societ`a Italiana
di Fisica.)
We have seen that the observed matter distribution in the universe follows some
sort of multiscaling behavior. If galaxies and clusters of galaxies with increasing richness are considered as different realizations of the selection of a density
threshold in the mass distribution, the multiscaling argument implies that the corresponding values of the correlation dimension D2 must decrease with increasing
density.
We shall show the correlation integral for galaxy samples and for cluster samples in the range [10, 50] h 1 Mpc. In this range of scales gg (r) does not follow a
power-law shape, while C (r) is nicely fitted to a power law. For galaxies we have
analyzed the CfA1 sample, the PiscesPerseus sample, and the QDOT-IRAS redshift survey. The cluster samples are the Abell and ACO catalogs, the Edinburgh
Durham redshift survey, the ROSAT X-ray-selected cluster sample, and the APM
cluster catalog.
153
2002 by CRC Press LLC
Multiscaling
141
Figure 4.19 The correlation integral for different galaxy and cluster samples in arbitrary
units. Objects corresponding to higher peaks of the density field provide lower slopes, i.e.,
lower values of the correlation dimension. (Reproduced, with permission, from Martnez
et al. 1995, Science, 269, 12451247; c 1995, American Association for the Advancement
of Science.)
Fig. 4.19 shows that three straight lines fit reasonably well the eight samples
analyzed. All the cluster samples have a correlation integral well fitted by a power
law with exponent D2 ' 2:1. A value of D2 ' 2:5 appears for the optical
galaxy catalogs: the CfA1 volume-limited sample and the PiscesPerseus survey
within the range [10, 50] h 1 Mpc. Finally, a value of D2 ' 2:8 is obtained for
the QDOT-IRAS galaxies. Note that there is no contradiction between the value
D2 ' 1:3 for the range scale [1, 10] h 1 Mpc provided at the end of Section
4.5 for the CfA1 sample and the value reported here. In fact, as we have already
noted in the discussion in Section 4.4.3, D2 for the galaxy distribution is a scaledependent quantity varying from 1.3 at small scales, where the approximation
D2 ' 3
is valid, to 3 at large scales where the distributions are much
more homogeneous. These results probe the multiscaling behavior of the matter
distribution in the universe. The fact that D2 for IRAS galaxies is larger than for
optical samples indicates that IRAS galaxies are less correlated than optical galaxies; this is nicely interpreted if optical galaxies correspond to higher peaks of the
density field. Clusters of galaxies have stronger correlations than single galaxies,
corresponding to the highest peaks of the background matter density.
154
2002 by CRC Press LLC
142
4.7 Lacunarity
Let us recall again the definition of the fractal massradius dimension D. If we
have a ball of radius R, centered on the point belonging to a fractal structure, then
the number of points in the ball (its mass) is given by
M (R) = F RD ;
(4.19)
where the prefactor F is a function (for deterministic fractals) or a random variable (for random fractals) that does not depend on R. Usually the main attention
is focused on the determination of the fractal dimension (mass dimension), but
that does not describe a fractal completely. Fractals of similar dimensions may
have a completely different appearance, as illustrated in Fig. 4.20. We compare
there two point distributions one is an observed galaxy distribution in a thin slice
from the Las Campanas survey (see Chapter 1) and another is its fractal model,
the RayleighLevy flight, introduced in Section 4.2.1.
The two point distributions have the same fractal dimension, as shown by the
comparison of the ln M ln R curves. But the overall appearance of the two distributions is different; they have different lacunarities. Lacunarity is a term that
has been used loosely to describe the properties of the prefactor F . Blumenfeld
and Mandelbrot (1997) proposed describing it by a series of variability factors
Sk =
Ck
;
C1k
(4.20)
where Ci is the i-th cumulant of the prefactor F . The simplest of these factors is
the second-order variability factor
E f(F F )2 g
:
(4.21)
F 2
The distribution of the prefactor F for one-dimensional Levy flight of dimension D is the MittagLeffler distribution with the density
1 ( 1)k 1
1X
sin(kD) (kD)xk 1
fD (x) =
k=1 (k 1)!
=
(Blumenfeld and Mandelbrot 1997). Using that distribution, Blumenfeld and Mandelbrot derived a simple expression for the variability factor:
(D) =
2 2 (1 + D)
(1 + 2D)
1:
For larger dimensions the distribution is not known, but the variability factor
is easy to estimate from observations. The variability factors (lacunarities) for
our two distributions are also given in Fig. 4.20 (the lower right panel). For a
strictly fractal distribution this factor should not depend on the radius of a ball
R and any such dependence is evidence for nonfractality. Fig. 4.20 shows that
155
2002 by CRC Press LLC
Lacunarity
143
10
3.5
3
2.5
ln M
6
2
1.5
-2
-4
1
0
3 4
ln R
3 4
ln R
Figure 4.20 Comparison of a Las Campanas survey slice (upper left panel) with its Levy
flight model (upper right panel). The fractal dimensions of both distributions coincide, as
shown by the ln M ln R curves in the lower left panel, but the lacunarity curves (in the
lower right panel) differ considerably. The solid lines describe the galaxy distribution, and
the dotted lines show the model results. The monotonous decrease of the lacunarity curve
for the LCRS slice indicates that the distribution is not as perfectly fractal as suggested by
the lower left panel.
both distributions are not strictly fractal. At small scales this is due to the Poisson
noise, as in the Poisson distribution
1
1
2 = 2;
n R
and n
is the mean density
144
157
2002 by CRC Press LLC
CHAPTER 5
146
= l LV = l s l :
159
2002 by CRC Press LLC
147
Fig. 5.2 shows a realization of a segment Cox process generated inside a cube
of side length L = 100 h 1 Mpc. The chosen values for the parameters are
s = 10 3 ; l = 0:6, and l = 10 h 1 Mpc. Point patterns containing N ' 6; 000
points are produced with this selection.
To calculate the correlation function of this point field we have to consider that
the driving random measure of the point field is equal to the random length measure of the system of segments. Stoyan, Kendall, and Mecke (1995) showed that
the two-point correlation function of the Cox point process equals the two-point
correlation function of the system of segments, which reads
1
1
(5.1)
2
2r LV 2rlLV
l and vanishes for larger r. It is worth noting that the expression (5.1)
Cox (r) =
for r
does not depend on the mean number of points per unit length of the segments.
The K function, the integral of the correlation function (see Eq. 3.45), is therefore:
8
<
4 r3 + r 2 r ;
r l;
s l
l
KCox (r) = 43
1
3
:
r > l;
3 r + s ;
and the correlation dimension D2 can be calculated analytically simply as:
r dK (r)
D2 (r) =
:
K (r) dr
(5.2)
(5.3)
148
Figure 5.2 Simulation of the segment Cox process with parameters s = 10 3 ; l = 0:6,
and l = 10 h 1 Mpc. N = 6; 007 points have been plotted. (Reproduced, with permission,
from Pons-Bordera et al. 1999, Astrophys. J., 523, 480491. AAS.)
In the plot the solid line of the bottom main panel represents the expected theoretical function given by Eq. 5.2. As we see, the empirical estimate of K (black dots)
reproduces the expected theoretical behavior quite satisfactorily. The edge correction included in the estimator has not destroyed the goodness of the estimates; in
particular, it has not introduced spurious homogenization. The estimator used (see
Eq. 3.50) works well not only in the easy case of absence of structure, which
represents Poisson processes (Martnez et al. 1998), but it has also been able to
reproduce quite exactly the very precise value of K for a clustered Cox process.
This test gives us enough confidence to believe that the K results obtained from
the galaxy samples effectively reflect the structure existing there.
Models related to these kinds of Cox processes have been used in cosmology to
mimic the galaxy distribution. For example, Buryak and Doroshkevich (1996) devised a model in which galaxies are randomly placed on straight lines and planes.
161
2002 by CRC Press LLC
149
Figure 5.3 Bottom panel: The average and rms error bars of the function K (r ) for 10
realizations of the Cox point processes (solid circles). The inset shows the two-point correlation function of this stochastic model. Top panel: The local correlation dimension D2
with rms errors calculated by means of a five-point weighted loglog least square fit on
the average of K . In both panels the solid line shows the theoretical values, while the dotted line in the bottom panel corresponds to KPois (r ). (Reproduced, with permission, from
Martnez et al. 1998, Mon. Not. R. Astr. Soc., 298, 12121222. Blackwell Science Ltd.)
150
can be said for the function 1 + (r). Looking at the top panel of Fig. 5.3, we
can see the behavior of the empirical local correlation dimension D2 calculated
over the average of the 10 realizations of the Cox processes together with the rms
deviations. The solid line represents the expected theoretical values (5.3). Again,
we can see the reliability of the estimates. But what is more interesting in this
example is the long plateau observed in the plot of the correlation dimension. The
value D2 ' 3
' 1 remains nearly constant for a broad range of scales, due
to the particular behavior of the K function for this model. After the fractal
behavior, a transition to homogeneity is clearly appreciated in both D2 and K (r).
At this point we want to remark that, in the same way that the term fractal is
not appropriate for the Cox process, even with a correlation function decaying as
a power law at short scales (Stoyan 1994), the galaxy distribution, even holding a
similar property, is not a fractal in a rigorous sense. However, in a looser use of
the term fractal (Avnir et al. 1998), it could be appropriate to talk about a fractal
regime to describe the range of scales where K (r) follows a power law, bearing in
mind that a real self-similar point pattern, for example the SoneiraPeebles model
described in Section 4.2.2, satisfies other conditions (self-similarity) apart from a
power-law decaying correlation function.
151
-1
1.5 h Mpc
zi
Figure 5.4 A Matern process is sketched here. The radius of the ball centered at each
parent point is R = 1:5 h 1 Mpc and = 5 is the mean number of points per cluster.
(Reproduced, with permission, from Kerscher et al. 1999, Astrophys. J., 513, 543548.
AAS.)
35
30
25
(r)
20
15
10
5
0
0
0.5
1.5
2.5
r (h-1 Mpc)
Figure 5.5 The correlation function of a Matern process with parameters = 0:01, = 5
and R = 1:5 h 1 Mpc.
(r) =
8
>
<
>
:
1 3
8R6 R
0;
r
2
2
2R + 2r
1
X
e n
;
n=2 (n 2)!
0 r < 2R;
r > 2R:
Van Lieshout and Baddeley (1996) have derived the formula of the J -function
corresponding to this point field, introduced in Section 3.8 In three dimensions
164
2002 by CRC Press LLC
152
JM (r) =
8
<
:
V (x;r;R) d3 x;
1
Vol(BR ) BR e
e ;
0r2R;
r > 2R;
(5.4)
where
V (x; r; R) =
Vol(Br (x) \ BR )
Vol(BR )
denotes the ratio of the volume of the intersection of two balls to the volume of
a single ball. Here Br (x) is a ball of radius r centered at the point x, while BR
is a ball of radius R centered at the origin. This quantity can be calculated from
basic geometric considerations (Stoyan and Stoyan 1994). In three dimensions,
the expression is
V (x; r; R) =
8
>
>
<
c3 x3 + c1 x + c0 + c 1 x 1 ;
>
>
:
r3 =R3 ;
1;
0r<R; R r<x<R;
or Rr2R; r R<x<r;
0r<R; 0xR r;
Rr2R; 0xr R:
c3 = 1 3 ;
c1 =
16R
3
c0 = 12 r 3 + 1 ; c 1 =
R
3 r2 + 1 ;
8 R3 R
3 2r2 r4 R :
16 R R3
Fig. 5.6 shows JM (r) for R = 1:5 h 1 Mpc and several values of . Obviously,
J (r) discriminates between the varying richness classes of the Matern cluster
processes.
Matern point processes can be easily modified by changing the distribution law
for locating the points within each cluster. For example, if points are scattered
within the balls following a symmetric normal distribution, we get a modified
Thomas process. This process has a known correlation function (Cox and Isham
1980) following an exponential law.
165
2002 by CRC Press LLC
153
Figure 5.6 The function JM for Matern processes with parameters R = 1:5 h 1 Mpc and
= 1; 3; 10; 30 (bending down successively). The solid lines correspond to the analytical
expression (5.4) while the areas represent the rms fluctuations of the J -function calculated
using 50 realizations of the point process. (Reproduced, with permission, from Kerscher et
al. 1999, Astrophys. J., 513, 543548. AAS.)
Barlett (1964) has shown the formal equivalence between Cox processes and
NeymanScott processes when, in this latter type, the number of points per cluster
follows a Poisson distribution.
5.3 The Voronoi model
The Voronoi tessellation (Voronoi 1908) is a partitioning of space uniquely defined by a discrete point set. Each point of the set (nucleus) is surrounded by a
Voronoi cell that encloses that part of space that is closer to its nucleus than to
any other nucleus. The Voronoi tessellation is quite dependent on the initial distribution of nuclei. Depending on the degree of clustering or spatial correlation
between points in the initial point set, the corresponding tessellation presents a
complete different aspect. Fig. 5.7 shows the Voronoi tessellation corresponding
to six different planar point distributions where different clustering patterns have
been simulated, from regular (anticorrelated) distributions to distributions with
166
2002 by CRC Press LLC
154
155
Figure 5.7 The Voronoi tessellation corresponding to six different point patterns (as nuclei
of the tessellation). The degree of clustering of the nuclei increases from top to bottom and
left to right. The greater the clustering, the greater the variability of the shapes and sizes
of the Voronoi cells. (Reproduced, with permission, from van de Weygaert 1991, Voids and
the Geometry of the Large Scale Structure, Leiden University.)
168
2002 by CRC Press LLC
156
Figure 5.8 Stereoscopic pair of three adjacent Voronoi cells. (Courtesy of Rien van de
Weygaert.)
157
1000
100
(r)
10
0.1
0.01
0.1
1
r (h-1 Mpc)
Figure 5.9 A point field based on the Voronoi model. The top panel shows the distribution of vertices of a Voronoi tessellation built from 1,500 randomly distributed nuclei. A
HammerAitoff projection is shown at the center, while the bottom panel shows the twopoint correlation function of this process, which follows a power law with exponent 2.
170
2002 by CRC Press LLC
158
Figure 5.10 Two realizations of point processes based on the scattering of points along
filaments (left panel) or walls (right panel) of a Voronoi tessellation.
idea comes from the observational fact, clearly noted in Chapter 1, that galaxies
cluster around filaments and walls. Bearing this in mind, we can use the Voronoi
foams as a skeleton with an evident cell-like structure, and place galaxies as desired:
1. At the edges of the tessellation (to populate filaments)
2. At the walls of the tessellation (to populate walls)
3. At the vertices of the tessellation (to populate clusters)
4. Or in any combination of the above elements with variable proportions
Different methods can be used to distribute the galaxies within a given Voronoi
element. The distribution can be uniform or follow a given prescription such as
a power-law density profile or a Gaussian distribution. Fig. 5.10 shows two realizations of this kind of Voronoi-based point process. In the left panel, points were
placed following a Gaussian distribution along the edges of a Voronoi tessellation
with Poisson-distributed nuclei, while in the right panel, points were placed populating the walls. This is obviously a highly interesting family of Voronoi-based
point processes that merits further investigation both analytically and by Monte
Carlo simulations.
171
2002 by CRC Press LLC
159
Figure 5.11 A planar point process consisting of 20 points is depicted together with its
Voronoi tessellation (left panel: solid lines) and its Delaunay tessellation (right panel:
solid lines). In the left panel, the shaded region marks the Voronoi cell of a given point,
while the right panel indicates the contiguous Voronoi cell associated with the same point.
(Reproduced, with permission, from Schaap and van de Weygaert 2000, Astr. Astrophys.,
363, L29L32.)
160
statistical distributions of discretely sampled velocity fields. The use of tessellations for spatial interpolation was anticipated by Sibson (1981). More recently,
Braun and Sambridge (1995) have applied similar methods for solving partial
differential equations.
5.4 Statistical models for the counts in cells
In Section 3.6 we introduced the statistics of the counts in cells, where P (N; V )
represented the probability that a randomly placed cell of volume V contains N
galaxies. Several models of galaxy clustering are based on particular choices of
this function, which are physically motivated. In some cases the physical motivation comes from the continuous version of the counts in cell statistics, the onepoint distribution function of the density field. In other cases, as for the Saslaw
distribution, more general physical principles are used to invoke such distributions.
5.4.1 The lognormal model
Hubble (1934), analyzing very preliminary projected catalogs of galaxies, was the
first to notice that the distribution of galaxy counts in two-dimensional cells could
be well approximated by a lognormal distribution.
The lognormal model for the density field in cosmology was introduced by
Coles and Jones (1991), who found how this random field could be obtained from
the gravitational evolution of an initially Gaussian field. They also present a model
for the distribution of galaxies based on the lognormal one-point probability distributed function of the smoothed field. Fig. 5.12 shows different realizations of
this model with different degrees of clustering obtained by increasing the variance
of the underlying Gaussian field. It is worth mentioning that a similar cluster process was introduced much more recently in the field of spatial statistics by Mller,
Syversveen, and Waagepetersen (1998). These authors termed this model the log
Gaussian Cox process and they present more sophisticated ways of simulating
point patterns of this model.
Assuming the Poisson model to extend the continuous lognormal distribution
to provide a discrete distribution of galaxies, Coles and Jones (1991) derived an
expression for the counts-in-cells probabilities
(ln(= ) )2
1 1 1 n 1
e exp
d;
P (N = n) = p
22
2 n! o
where = is the intensity of the Poisson process, represents the mass density
(the continuous density field), and is a normalizing constant introduced to match
the correct number density of objects. The correlation function of this model is
presented in Section 7.5.
173
2002 by CRC Press LLC
161
Figure 5.12 Four realizations of a point process based on a lognormal random field with
underlying Gaussian fluctuations having variances (a) 0.01, (b) 0.1, (c) 1.0, and (d) 10.0.
(Reproduced, with permission, from Coles and Jones 1991, Mon. Not. R. Astr. Soc., 248,
113. Blackwell Science Ltd.)
162
(Saslaw 2000). This function was derived assuming that a system of gravitating
point masses in an expanding universe is in thermodynamical equilibrium. The
probability of finding N galaxies in a randomly placed volume V is
N (1 b)
N (1 b) + Nb N 1 e [N (1 b)+Nb] :
(5.5)
N!
= n V is the expectation value of N (n is the mean galaxy number
Here N
density), and the parameter b is defined as
U
b=
;
2K
where U is the average gravitational energy of density fluctuations and K is the
P (N; V ) =
average energy of peculiar velocities (fluctuations from the uniform Hubble expansion). As the distribution function is derived in a thermodynamical framework,
an important quantity is the average temperature T of the point ensemble, defined
by
mN hv2 i 3
= NT;
2
2
where m is the mass of a particle and hv 2 i is the dispersion of peculiar velocities.
The energy ratio b can now be written as
Z
2Gm2 n (n; T; r) 2
r dr;
b=
(5.6)
3T
r
V
where G is the gravitational constant and (r) is the correlation function of denK=
sity fluctuations. Eq. (5.6) shows that the correlation function depends both on
the mean number density and the temperature of the ensemble. Saslaw (2000) has
adopted the dependence
b=
b0 n T 3
:
1 + b0 n T 3
h(N )2V i =
N
:
1 b(V ) 2
175
2002 by CRC Press LLC
163
Assuming adiabatic (slow) evolution of structure, Saslaw derives the time evolution of b in cosmological context:
b1=8
a(t) = a?
;
(1 b)7=8
where a(t) is the dimensionless scale factor and a? is a constant defined by the
initial value of b at a certain epoch t.
Assuming that the kinetic energy of the fluctuations is proportional to their
potential energy (gravitational quasi-equilibrium of the same degree on all scales)
Saslaw obtains the peculiar velocity distribution density
2
2
22 (1 b)
2 u 1
(1
b
)
+
bu
e (1 b) u u du:
(u2 + 1)
Here the normalized peculiar velocity u = v (R=Gm)1=2 , R is the size of the
ensemble volume, is the gamma function, = hu2 i is the (normalized) peculiar velocity dispersion, and the density form factor = hR=riP hu2 i 1 , where
hR=riP is the value of R=r averaged over a Poisson distribution with mean num.
ber density n
f (u)du =
Specifying both the velocity and number density distributions gives a complete
description of a gravitating system, but the velocity distribution also is important
for statistical studies. We shall see in subsequent chapters that peculiar velocities
can severely distort our estimates of the observed density distribution.
Borgani (1993) has shown that this model exhibits a very interesting multifractal behavior. From the expression of the P (N; V ) it is rather straightforward to
show that the Renyi dimensions are
(
D(q) =
3 1
q q ;
3
;
q 1=2;
q > 1=2;
176
2002 by CRC Press LLC
(5.7)
CHAPTER 6
Formation of structure
6.1 Introduction
Because our observations are based on the properties of different physical fields,
objects, and their populations (CMB, galaxies, etc.), we must understand how
these objects form and evolve and how their properties are connected with those
of the universe and with one another.
The present consensus in cosmology is that the observed structure developed
from small initial perturbations of the physical fields (density, velocity, gravitational potential, etc.) resulting from the instability of the Friedmann models for
small perturbations. The Newtonian theory of gravity also predicts the instability of a self-gravitating distribution of matter. While in the Newtonian theory the
growth rate of perturbations is exponential in time, in cosmology the expansion of
the universe slows the growth of perturbations and the growth rate is typically a
power law. This slow growth continues until local densities become large enough
to stop the expansion. After that, structure starts to evolve rapidly, forming the
objects we observe now.
Below we delineate the minimal theory of formation of structure needed to
understand the present statistical applications. A detailed representation of the
theory can be found in Peebles (1980).
6.2 Dynamics of structure
Using the theory of the initial inflational stages of the universe, it can be shown
that the present structure evolved from small fluctuations of the continuous matter
density, velocity, and temperature fields. It can be also shown that when we are
describing physics at spatial scales substantially less than the radius of the curvature of the universe, we can use the usual Eulerian hydrodynamics and Newtonian
mechanics. The equations describing the evolution of the matter are then (see Peebles 1980; Peacock 1999)
D
+ r (w) = 0;
Dt
Dw
1 rp r;
=
Dt
= 4G;
177
2002 by CRC Press LLC
(6.1)
166
Formation of structure
where D=Dt
@=@t w r is the convective derivative, is the density of
matter, p is the pressure, w is the velocity field, is the gravitational potential,
is the Laplacian, and G is the gravitational constant. The first equation is the
continuity equation, the second is the equation of motion, and the third is the
Poisson equation. These equations have to be complemented by the equation of
p and, because this equation usually depends on temperature, an
state p
equation of energy, as well. During these stages of evolution when we need to
use dynamics to predict statistical descriptors the pressure term in the equation of
motion is negligible, and we shall omit it from further discussion. Pressure effects
are important for the very early universe (radiation pressure) and at final stages
of formation of objects when they counteract gravitational forces. There are no
statistical applications for the first case yet and the pressure effects for the second
case are accounted for in statistical theories in an approximate manner.
In cosmology it is more convenient to work in comoving coordinates x that are
connected to the Eulerian coordinates r by
= ()
r = a(t)x;
(6.2)
where a
we get
= ()
D
+ r (u) = 0:
Dt
(6.4)
We can use the transformation (6.2, 6.3) to transform the Eulerian equation of
motion into comoving coordinates, but it is easier to derive it as Peacock (1999)
has done, differentiating x in (6.2) twice to get
D2 x
Dt2
The last equality comes from the Eulerian equation of motion, and we have used
the subscript r to show that this gradient is calculated with respect to the Eulerian
=a rx . Using
coordinates. The transformation formula (6.2) says that rr
the Eulerian equation of motion together with the Poisson equation for the case
a=a r, we get
of the smooth motion u 0, w
= (1 )
= (_ )
Dw
+ w rw = aa x = r ;
Dt
r
178
2002 by CRC Press LLC
Dynamics of structure
167
0 = 4G( ):
r
There is no trouble with this Poisson equation because the volume integral of the
source term, the fluctuating density, is zero.
We shall drop below the coordinate index r and the prime and write the fluctuating potential as . We will also use the density contrast instead of density:
(x) =
=
(x)
:
Note that
C , if calculated in comoving coordinates, does not depend on
time, and is related to the Eulerian E we have used thus far as E a 3 C . The
equations of dynamics in comoving coordinates now become
=
D
+ r [(1 + )u] = 0;
Dt
a_
1
Du
+
2
u = 2 r;
Dt
a
a
= 4Ga 1:
(6.5)
(6.6)
(6.7)
Using the comoving density simplified the derivation of the continuity equation. You will not see these equations often in the cosmological literature, however, because the Eulerian (physical) density field is preferred in the dynamical
equations. Because the density contrast is defined as a normalized fractional amplitude, its value is the same in both cases. The mass of a volume element is
dM = (1 + ) dV = (1 + ) dV ;
and because dV = dV , the density contrasts are really the same. The
E
@ 1
+ r [(1 + )v] = 0;
@t a
@v 1
+ v rv + aa_ v = a1 r;
@t a
179
(6.8)
(6.9)
168
Formation of structure
= 4Ga2 :
(6.10)
These equations are approximate only in the sense that we cannot use them for
scales close to the curvature radius and in extremely strong fields (e.g., in black
holes), but everywhere else they are perfectly applicable. In particular, we can use
them to describe the large densities and velocities that occur in the observable
universe.
6.3 The linear approximation
6.3.1 Density evolution
Although the equations of dynamics do not appear too complex, they are impossible to solve analytically, in general, but nonetheless many useful approximations
have been found. In cosmology even the most simple linear approximation goes a
long way the amplitude of the density fluctuations after recombination, when
3 . We shall also see later that
the real growth of the structure starts, is about
at many observationally relevant scales the density contrast is still so small that it
can be described by linear dynamics.
In the linear approximation we can replace convective derivatives by partial
time derivatives, as juj . We can also neglect the term r u in the continuity
equation and write
10
@
+ r u = 0:
@t
(6.11)
@2
+ r @@tu = 0;
@t2
r @@tu + 2 aa_ r u = a12 :
(6.12)
(6.13)
Combining now equations (6.12, 6.13, 6.11) and (6.10) we get a second-order
equation for the evolution of the density contrast:
a_
+ 2 _
a
4G = 0:
(6.14)
Because this equation does not depend explicitly on the spatial coordinates and
thus is local, we have replaced partial time derivatives with ordinary ones. Although gravity is a nonlocal force, it enters in the linear approximation only
through the Laplacian of the potential, which is local. This property disappears
in subsequent approximations, which are nonlocal.
Equation (6.14) is a linear second-order ordinary differential equation, and its
solution can be written as
169
()
()
The partial solutions D1 t and D2 t are called, respectively, the growing and
the decaying modes (obviously, we can order them in this way).
We shall illustrate the solutions in a simple case of the Einsteinde Sitter ( M
) universe. In this case
= t2 , a=a
= t , and the density
;
contrast equation is
4 = 2 (3 ) _ = 2 (3 )
4
2
+ _
3t 3t2 = 0:
1
=0
This equation is homogeneous in t; thus, the partial solutions are power laws,
s
. Substituting this form into the equation we get two solutions s
= and s
, the first for the growing mode D1 t
t2=3 and the second for the decaying
t 1 . The decaying mode is usually discarded in theoretical studies;
mode D2 t
for longer periods of evolution its amplitude becomes negligible compared with
that of the growing mode if we do not have very specific initial conditions.
As we see, the growing mode is proportional to a t in this simple model. We
see also that although the density contrast is growing with time, the physical
actually gets smaller, e a 2 t 4=3 ; in the
overdensity e
linear regime the infall of matter into objects is slower than the expansion of the
universe. The formation of structure is due to the fact that at some limiting value
of the process becomes nonlinear and causes a really fast growth of density.
. Of
Another simple case is that of the empty universe, when a t t and
course, there can be no matter density perturbations without matter, but as before,
we can consider this model as a good and simple approximation of a low-density
universe. The linear density contrast equation is
=2 3
()=
( )=
()
=
()
= 0
+ _ = 0:
t
()=
This equation is easy to solve and it gives for the modes the results D1 t
const
t 1 . If we recall the evolution of the instantaneous cosmological
and D2 t
parameters in the case of an open model (it is close to the flat case until a certain
redshift and then starts to expand faster), we can say that the growth of the density
contrast freezes in an open model at a certain epoch.
In a general case the linear density contrast equation (6.14) can be solved almost completely analytically (Saar 1973). Changing the argument from time to
redshift by
()=
dz
dt
H0 (1 + z )E (z )
E0
00 +
E
=1+
1 0 3
x
2
x
= 0:
E2
Here x
z is a redshift-type variable and primes denote differentiation with
respect to x. Defining now a new function x by
()
= E
181
170
Formation of structure
E 00
E 02 E 0 3
x
+ E 00 +
= 0:
x
E
x 2 E
It can be checked that the coefficient for is zero for any E 2 (x) that is a cubic polynomial in x, and that this class also includes the E 2 for the canonical
(
;
) models; see (2.21). Thus we can write
E 0
00
0
E + 3E
= 0;
x
E 00 +
3E 0
which gives
x
0 = A 3 ;
E
(z ) = AD1 (z ) + BD2 (z );
D1 (z ) = E (z )
and the decaying mode is
Z
z
1 (1 + z 0 )
dz 0
E 3 (z 0 )
D2 (z ) = E (z ):
= 0 models
The integral for the growing mode can be taken analytically for all
(Sahni and Coles 1995):
1 + 2
+ 3
D1 (z ) =
j1
j2
M
where
I (
8
>
>
<
)=>
>
:
+ 3
(1 +j1z)
(
j5z2+ 1) I (
1 log pp
z + 1 + pp1
;
z + 1 r1
arctan
z +11 ;
; z );
(6.15)
< 1;
> 1:
For the models with a cosmological constant the solutions have to be found
numerically.
6.3.2 Velocity evolution
Although we had to eliminate the velocity field from the dynamical equations in
order to get the equation for the density contrast, we must also describe it, as it is,
in principle, directly observable.
The linear velocity equation is
@u
+ 2 aa_ u = g;
@t
182
(6.16)
171
where the gravitation acceleration g can be written, solving the Poisson equation,
as
Z
0
g(x) = Gr
(x )
jx0
xj
(6.17)
The velocity field, as any vector field, can be written as the sum of a longitudinal
(nonrotational) field and a transverse (rotational) field, u uL uT , ruL 0,
r uT . Because the nonhomogeneous term in (6.16) is a gradient of a scalar
function, the linear velocity equation splits into two:
= +
=0
@u
@t
@u
@t
+ 2 aa_ u = 0;
+ 2 aa_ u = g
T
uT
= C (x)a 2(t);
T
where the constant field CT is purely rotational; initial rotational motions decay
rapidly.
The equation for the longitudinal (divergent) part can be solved, noting that
both the velocity (6.19) and the acceleration (6.17) are proportional to the density contrast . Because this is a sum of two modes and our equations are linear,
uL also must be a sum of two modes, linear in density contrast. Comparing the
velocity equation (6.16) and the density equation (6.14) it is easy to see that the
velocity component resulting from the first density mode D1 t can be written as
u=
()
dD1
4GD1 dt
g
(6.18)
(the normalization of g is chosen to eliminate its dependence on time). Substituting (6.18) into the velocity equation (6.16) we get the equation for the density contrast that is solved by D1 t , so our solution (6.18) must be correct. Just as there
are two modes for the density contrast, there are two modes for the longitudinal
velocity: one growing, the other decaying. Recalling the results for D1 t ; D2 t
from above, we can write for the standard Einsteinde Sitter model:
()
()
D1 (t) t2
D2 (t) t
()
3 ; u t 1=3 ; v t1=3 ;
L1
L1
1 ; u t 2 ; v t 4=3 :
L2
L2
We see that the comoving velocities diminish in time, and the word growing
mode is justified only for peculiar velocities in the physical space. The minus
sign for the second mode means that these velocities are directed against the gravitational acceleration, working to erase the initial density contrast.
; ), we get
For the second example, the empty universe ( M
=0
=0
v 1 = 0;
D1 (t) = const; u 1 = 0;
D2 (t) t 1 ; u 2 t 2 ; v 2 t 4 3
L
183
2002 by CRC Press LLC
172
Formation of structure
(zero velocities mean freezing). The growing and decaying modes for density and
velocity are connected. The usual approach is to discard the decaying modes and
use only the growing mode. In a generic situation the amplitudes of both modes
should be similar at the start of the growth of structure (after recombination, at
z
). Because the ratio of the amplitudes D2 =D1 t 5=3
z 5=2
8 , so
(in the Einsteinde Sitter universe), this ratio for the recent past is about
we can safely disregard the decaying mode.
1300
(1 + )
10
d
dt
d
log = H d log D1
= da
= a_ a dd log
dt da
a
d log a
(if we neglect the decaying mode, we know the time behavior of the density contrast, a D1 a ). The continuity equation (6.11) now gives us
()
()
1 r u:
Hf (a)
(6.19)
For direct comparison with observations we would use the peculiar velocity v and
compute the divergence in physical coordinates. The scaling (6.2) ensures that the
above formula remains the same (rr v rx u).
This expression is frequently used in applications. The function f is called the
dimensionless growth rate. It can be computed, using the above expressions for
D1 , but in most applications approximate expressions are used. For the present
a good approximation is
epoch and in the case of
=0
f (
; z = 0) =
M
d log D1
j
0 6 :
d log a =0
:
M
(6.20)
f (
;
)
0 6 +
70
:
M
1 +
2
They also found that the approximation (6.20) can be used for all epochs z , if we
use the instantaneous value of the parameter M z (formula 2.27):
()
f (
; z )
0 6 (z ):
:
M
184
2002 by CRC Press LLC
173
The formula for the velocity (6.18) can be rewritten, replacing the time variable
by a and using the definition of the function f (6.20):
v=
H0 f
2f g :
g=
4G 3H0
(6.21)
(2.18).
r referred to the physical space, the coordinate x to the comoving space). In both
spaces mass points (objects) change their coordinates, if their velocities are not
zero. An alternative description is the purely Lagrangian one, where the main
coordinates are the labels of the particles (e.g., these labels could easily be their
Eulerian positions at a freely chosen moment). The dynamics is now described by
the mapping from the Lagrangian to the Eulerian coordinates, which defines all
other physical fields.
The reader can find a strict mathematical representation of the cosmological Lagrangian dynamics in Buchert (1992) and a more pedestrian approach in Bouchet
et al. (1995).
We shall start by taking another look at the formula connecting velocity and
acceleration formula (6.21) in the linear regime. Besides showing how we can
calculate velocity given the acceleration, it also implies that the direction of the
velocity is constant; particles move along straight lines, otherwise the directions
of the velocity and acceleration would differ.
If we know that the growing mode of the velocity is irrotational, we can introduce the initial velocity potential q and write for the linear approximation in
the Lagrangian approach
( )
(6.22)
()
()
= 0 jx
i;k
j 1 = 0 j + D1 (t) ; j 1 ;
ik
ik
=
=
where jxi;k j is the Jacobian of the mapping (6.22), and we use the notation x;i
@x=@qi for partial derivatives. Let the eigenvalues of the deformation tensor fik
;ik be i ; i
; ; these are the contraction (or expansion) coefficients
along the three eigenvectors (fik is a symmetric tensor and can be easily diagonalized). The formula for density simplifies then to
=123
(6.23)
174
Formation of structure
(t) 1) we get
(t) = D1 (t)(1 + 2 + 3 );
thus the density contrast grows as it should in the linear regime (wherever the sum
of the eigenvalues of the deformation tensor is negative).
Differentiating the mapping formula (6.22) we get the comoving velocity
u=
dD1 (t)
r (q):
dt
= 4Ga2 D1 :
Replacing
= r=a2, we find
Thus, the velocity potential and the gravitational potential are proportional and
carry the same information. The label inertial motion that we used above to
describe the linear motion is somewhat misleading the initial velocities and
the initial gravitational field are closely connected, thus we may also say that the
motion is purely gravitational. This is the result of having chosen the growing
; ) the denmode of fluctuations. In the Einsteinde Sitter model ( M
sity contrast is proportional to the scale factor, the coefficient before the velocity
potential is constant, and both potentials remain constant in the linear regime.
The Zeldovich approximation goes farther than the usual Eulerian linear approximation. As proposed by Zeldovich (1970), it can be applied after the linear
regime, when the typical perturbation amplitude D1 t i and the density
contrast > . While in the linear regime it was the sum of the eigenvalues of the
deformation tensor that determined the amplitude of the density contrast, at other
times that amplitude will be determined only by the most negative eigenvalue of
the deformation tensor. If we write that eigenvalue as and designate the eigenvector to which it corresponds as e1 , we see that the density contrast grows along
e1 faster than in any other directions,
D1 t 1 , and it reaches infinity at
1 . Thus, the Zeldovich ansatz predicts the collapse of matthe moment D1 t
ter into planes (sheets). As the initial configuration becomes increasingly flatter
(more sheetlike) during the density evolution, the approximation works increasingly better. The reason is that for the collapse of plane-parallel configurations
of matter the acceleration is exactly constant it is perpendicular to the plane
of the symmetry and does not vary in time, because it is determined by the mass
coordinate of a point (the Lagrangian q in our case) only.
Comparison with numerical simulations shows that the Zeldovich approximation works well beyond the linear regime. It breaks down, obviously, when the
=1
=0
()
( )=
(1
186
2002 by CRC Press LLC
Exact solutions
175
trajectories of the particles begin to cross one another, and it does not describe
well spherical condensations, where the collapse is essentially three-dimensional.
The Zeldovich approximation is widely used in cosmological studies, both for
analytical description of large-scale motions and for setting up initial conditions
for numerical modeling we need to generate only one scalar function q to
calculate all the necessary initial fields.
( )
= cos( )
[ 2 2]
=2
(t=t0 )2 3 L2 k sin(kq);
2
13 2
u
3t0 (t=t0 ) L k sin(kq);
1
=
2
3
1 (t=t0 ) L2 k2 cos(kq) :
This sheet will collapse at t = t0 (Lk ) 3 = (2 ) 3 t0 .
We illustrate this simple solution in Fig. 6.1. We have chosen a moment rather
close to the start of the collapse t = 0:9t . The left panel shows the velocity potenx
=
=
tial, the velocity, and the density contrast in the Lagrangian coordinates, and the
right panel shows the same in the comoving (Eulerian) coordinates. The velocity
and the potential are arbitrarily rescaled. While there is not much happening in the
187
176
Formation of structure
14
12
10
8
6
4
2
0
-2
-4
14
12
10
8
6
4
2
0
-2
-4
-0.4 -0.2
0 0.2 0.4
q/L
-0.4 -0.2
0
x/L
0.2 0.4
Figure 6.1 Gravitational evolution of a plane wave in the Einsteinde Sitter universe. The
physical fields are shown at the time t = 0:9tc , where tc is the collapse time; the velocity and the velocity potential have been rescaled. The left panel shows the coordinate
profiles of the velocity potential , the comoving velocity u, and the density contrast
in the Lagrangian coordinates; the right panel shows the same in the comoving Eulerian
coordinates.
Lagrangian space apart from the growth of density, the coordinated movement of
matter is creating steep profiles in the Eulerian space. Note, however, that the amplitudes are the same in both panels. The high density contrasts that occur during
the evolution of structure are typical for the gravitational instability picture.
Another typical feature of gravitational collapse is the way the density evolves.
Fig. 6.2 shows the evolution of the central density of the collapsing plane (again
in the Einsteinde Sitter cosmology). As shown, the density contrast evolves for a
long time, almost up to , by the linear rule t2=3 . After that it starts growing extremely fast, becoming infinite in a finite time. In physics such a growth is
called explosive. This justifies the usual approximation of describing collapse
by the linear and highly nonlinear stages. As density contrasts are very high in the
nonlinear stage, the expansion of the universe no longer influences the dynamics
and the nonlinear stage can be treated as Newtonian.
The Zeldovich approximation predicts that the particles cross the central plane
tc with the same velocity at which they arrived there (the
of the sheet after t
velocity changes very slowly, as D1 t , compared to the extremely rapid growth
of the density at collapse). Although this crossing may seem strange, it is possible
and natural, as the main constituent of matter, the dark matter, can be thought of
as dust, tiny particles that pass one another without colliding. This sort of matter
will develop streams of different velocities that coexist at the same point of space.
Thus far, the Zeldovich model is correct, but the accelerations and the velocities
it predicts after collapse are incorrect, because it cannot account for the gravitational forces from the collapsed sheet itself. The further evolution of the sheet has
_ ()
188
2002 by CRC Press LLC
Exact solutions
177
1000
100
10
0.1
0.01
0.001
0.01
0.1
t/tc
Figure 6.2 The growth of the central density contrast of a plane wave in the Einstein
de Sitter universe. After a long linear growth period the density contrast becomes infinite
in a finite time tc .
been followed numerically in many papers, partly because it is one of the easiest
problems to understand and partly because the collapse of a plane is an excellent
test of a numerical code.
Similar flows can develop in the baryon component of the matter, too, if it is
condensed into compact objects. For example, stars in galaxies usually participate
in different flows that cross one another easily. In the early universe, however,
baryons are more likely to form a gas, and that gas will develop high pressure
in the high density regions, high temperatures, and shock waves that accompany
collapse.
6.4.2 Spherical collapse
There are two more exact solutions the collapse of a constant density sphere
and that of a similar ellipsoid. The ellipsoid solution is rather complicated and
is not used much, but the spherical collapse is a well-loved workhorse, and that
solution is especially popular in statistical applications. This is understandable,
because the simplest approximation for an isolated object is a sphere.
universe the dynamics of a gravitating sphere of
In a matter-dominated
a mass M and a radius R can be described by the Newtonian equation (we shall
work in physical coordinates for a while)
=0
d2 R
dt2
GM
:
R2
189
2002 by CRC Press LLC
178
Formation of structure
+ C;
(6.24)
where C is the constant of integration that determines the total energy of the
sphere (if C > , the sphere will expand forever; if C < , expansion will change
to contraction at a certain R). As we see, this equation is practically the same as
the Friedmann equation (2.15), because any spherical region in a universe expands
at the same rate as the overall expansion factor. This allows us to fix the value of
the constant C as C kc2 , where c is the velocity of light.
We are interested in the formation of an object, so we shall choose the case
C < ;k
. If we change now to a new nondimensional variable by
0 = 1
dt=d = R=c;
we can write
dR 2 2R
= R
d
R2 ;
(6.25)
(6.26)
where Rc GM=c2 .
As the reader can check, this equation has a (well-known) solution, the cycloid
R=R
(1 cos ):
(6.27)
t = t (
sin );
where
(6.28)
= R =c:
c
=0
(6.29)
t
which gives
t 3
2
6 1 20
"
1 6t 2 3
t 1+
6
20 t
t 3
R
R 2
c
1 12
190
2002 by CRC Press LLC
2
(6.30)
Exact solutions
179
3M 1 1 + 3 6t 2 3
=
4R3 6Gt2
20 t
=
ES
1 ;
= 6Gt
2
(6.31)
(6.32)
is the density of the Einsteinde Sitter model (2.26), so the (linear) density contrast of our sphere in that model is
2 3
6t :
3
=
20 t
=
(6.33)
As we know, all other models approach that model in the past, so their mean
model plus a small
density also can be written as the density of the M
density contrast, and the real density contrast of our sphere in any model is the
difference of those two. Another model-dependent quantity is the growing mode
of the density contrast D1 t . Because the final results do not depend much on
case,
the background model, we shall consider below only the simplest M
where (6.33) gives the true density contrast.
The meaning of the above exercise was to find a connection between the real
dynamics of the sphere given by the cycloid formulae (6.27, 6.28) and the linear
approximation.
The exact formulae tell us that the sphere starts to expand from a zero radius
until a maximum radius, from where it contracts to the zero radius again. The first
essential moment of the evolution is called turnover until that the sphere expands together with the universe, and after that real collapse begins. This moment
, the time is then tT tc and the radius of the sphere is
corresponds to
RT
Rc . The density of the sphere is
M= RT3 , and the ratio of that to
the background density is
=1
()
=2
=1
=
=3 4
ES
= 916 5:552
2
(to get that one has to use (6.29)). The density contrast at that moment is, accordingly, T : . It is also useful to know the density contrast for that moment
=
2
that would be predicted by the linear theory. By (6.33) it is Tlin
: .
4 552
= (3 20)(6 )
1 062
191
2002 by CRC Press LLC
180
Formation of structure
Moment
linear
Turnover:
Collapse:
Virialization:
4.552
1.062
1.686
1.686
177.7
lin
F
U + 2K = 0;
(6.34)
where U is the potential and K is the kinetic energy. At the point of maximum
, and the total energy is conserved, thus E UT
expansion (turnover) KT
UV KV . From (6.34) we get the initial kinetic energy KV
UV = , leading
to UT UV = . As the gravitational energy
=0
m(r)(r) 2
r dr;
r
0
(m(r) is the mass inside a sphere with a radius r) is inversely proportional to R,
we get R = R =2. The density of the virialized object will thus be 8 times
U
= 4G
higher than at the turnover, and as the time until virialization is twice as long as
that until turnover and ES t 2 , the relative density of the virialized object will
T =ES 2 : . The collapse density contrasts
be V =ES
are summarized in Table 6.1.
The specific numbers in that table refer to the Einsteinde Sitter model. Peacock (1999) derives an approximate expression
=8 4
=0
=2 9
177 7
1 + 178
vir
0:7
for
models, and Lahav et al. (1991) show that the value of influences
little the dynamics of the spherical collapse. The commonly used virial contrast
values (real and linear) in cosmological statistics are those from Table 6.1.
Fig. 6.3 illustrates the spherical collapse. The left panel shows the evolution
of the densities, both for the sphere and for the background. The density of the
sphere follows rather closely the background density almost until the turnover. At
tc ) the density of the sphere becomes infinite, but
the moment of collapse (t
=2
192
2002 by CRC Press LLC
5
4
3
2
1
0
-1
-2
-3
181
6
5
4
log()
log()
Exact solutions
sphere
3
2
exact
1
0
background
linear
-1
0.1
1
t/tc
10
0.1
1
t/tc
10
Figure 6.3 Collapse of a sphere in the Einsteinde Sitter universe. The left panel shows
the evolution of the density of the sphere and of the background density, the right panel
shows the behavior of the exact and linear density contrasts. The time unit tc is defined in
the text; the density units are rescaled.
stabilizes then at the level of 8 times the minimum (turnover) density. The density
contrast, shown in the right panel, grows continuously, and follows the t2 law
after virialization (the density of the sphere is constant, but that of the universe
decreases). We can also see the extremely fast density evolution at the late stages
of collapse, similar to that in the case of the plane collapse above (faster, in fact).
This figure also shows why we searched for the linear theory connection. The
point is that it can be used to specify the initial conditions for a collapse in the
past. For example, let us suppose that a certain class of objects is so young that
it must have formed in the immediate past. Setting then the present time for the
collapse epoch, we can predict the amplitude of the density contrast as
D (t)
(t) = 1:686 1
D1 (T )
= 11:+686z ;
where T is the age of the universe and the last equality is true for the Einstein
de Sitter universe. For any other hypothesis for the collapse epoch tc we have only
to replace D1 T by D1 tc in the above formula. As seen from the figure, this
prediction is correct almost up to the present time.
The spherical collapse formulae above do not refer to the density of the sphere;
the only variables are its radius and mass. Thus they can be used to describe the
collapse of spherical distributions of varying density profile. As the mass of the
sphere M is supposed to be constant, the different spherical shells should not
overtake one another. In cosmological applications usually d=dR < and this
condition is satisfied.
( )
( )
193
2002 by CRC Press LLC
182
Formation of structure
10
()
a_
a
u_ + 2 u = g;
194
2002 by CRC Press LLC
Numerical experiments
183
x_ = u:
The most time-consuming part is the calculation of the gravitational acceleration
g, as its value depends on the positions of all particles.
The first models found accelerations using the pairwise force summation over
all particles,
gi
=N
X
=0
xj
jx
xi
;
xi j3
(6.35)
where the indices i; j label particles. (In practice the force law is slightly modified
to account for the finite size of the particle; this is called force softening.) It is
evident, however, that this approach can be used only for rather small numbers
of particles N, because the number of operations needed to find all accelerations
scales as N 2 .
The number of particles necessary to model a representative volume of space
and to have a proper mass resolution is large. If we want to model a typical observed volume of (100 Mpc)3 , and we recall that the mean density of the universe
is : 11 M h2 M =Mpc3 , we see that by choosing 6 particles the mass of
a particle would be about 11 M , typical for a galaxy. So that huge model of
a million particles is still rather rough, able to explain formation of clusters and
larger structures, but not of galaxies themselves.
Pairwise force summation for 6 particles gives the operation count of about
13 . This is prohibitively large; thus, pairwise summation is not a useful method
for cosmological simulations.
Different algorithms for building numerical models of structure differ mainly in
the methods used to calculate gravitational forces. The first category is formed by
tree-codes that divide the computational volume into a hierarchy of cells until
the lowest level cells contain only one particle. When calculating the acceleration
for a particle, it is first determined at what angle a cell is seen from the position
of that particle. If that value is smaller than a given limit (the usual choice is
s=d < , where d is the distance to the cell and s is the size of the cell), the sum
of accelerations caused by the particles belonging to the cell is approximated by
the acceleration from their total mass at the mass center of the cell. This trick
N . A good description of the tree
reduces the total operations count to O N
code can be found in Hernquist (1987). Tree codes can be parallelized, and the
record particle number used is over 16 million.
Another approach, the particle-mesh method (PM), calculates accelerations
by first estimating the density on a fixed spatial grid (mesh), assuming the particle
size and its density distribution. The density distribution is assumed to be periodic
on the scale of the computational cube, allowing the Poisson equation to be solved
on the grid by the Fast Fourier Transform method (FFT). This is the step that
N in this case, too. The accelerations
reduces the operation count to O N
are calculated then by interpolating the gravitational potential. This method is
2 8 10
10
3 10
10
10
( log )
( log )
195
184
Formation of structure
Figure 6.4 Evolved density distribution from a N-body model. The point set has been uniformly diluted and only a few percent of the total N = 643 points are shown. (Reproduced,
with permission, from Pons-Bordera et al. 1999, Astrophys. J., 523, 480491. AAS.)
well described in Hockney and Eastwood (1988). As interpolation damps smallscale forces, an improved method, called P3 M, uses the FFT solver for large-scale
matter distribution and adds pairwise summed forces from neighboring points.
The application of that method to cosmology is demonstrated in Efstathiou et al.
(1985).
Fig. 6.4 shows an example of an evolved density distribution obtained this way.
It was generated using a P3 M solver for 3 points (a diluted version of the point
distribution is shown, otherwise the projection effects would destroy the picture).
The structures shown are typical for dark matter simulations and resemble, to a
large extent, the observed large-scale structure as well.
64
Numerical experiments
185
Figure 6.5 Thin slices of the spatial distribution of dark matter and galaxies from a recent
numerical simulation. The size of the region is 85 h 1 Mpc; the thickness of the slice is
8 h 1 Mpc. The top left panel shows the distribution of dark matter. In the other three
panels, galaxies are added as white circles. The top right panel shows the locations of
bright galaxies, and lower luminosity galaxies in the bottom panels are divided into two
groups those with a low star formation rate (bottom left panel) and those with a high
star formation rate (bottom right panel). (Reproduced, with permission, from Kauffmann
et al. 1999, Mon. Not. R. Astr. Soc., 303, 188206. Blackwell Science Ltd.)
186
Formation of structure
for pairwise force summation is kept approximately constant, however high the
local density grows.
Most of the codes described above are freeware. Their availability has provided
a strong impetus for the use of numerical models in cosmological studies. Both
Hydra and the adaptive P3 M (without gas dynamics) can be obtained from the
Hydra Consortium Web page (Hydra).
Fig. 6.5 shows slices from a recent simulation, an example of a typical numerical model of the structure. The basic simulation was done by the Virgo
Consortium, a group of astronomers pooling computational resources to create
detailed structural models based on the parallel version of Hydra (Jenkins et al.
3 particles, but no gas. The cosmology is specified by
1998). This model has
;
, and it is supposed that galaxies started to form later than
M
usual due to a neutrino background (the CDM model). The size of the cube is
h 1 Mpc, and the mass of a particle is 10 h 1 M : (Including gas dynamics
slows the computation considerably, and the best gasdark matter models com3 of both types of particles.) The dark matter distribution
puted thus far have
in a thin slice with thickness of h 1 Mpc (more than 1.5 million particles) is
shown in the top left panel of Fig. 6.5.
In order to predict the formation of galaxies in a dark matter model, the socalled semianalytical methods are frequently used recipes for the collapse of
gaseous objects, of their further merging, of star generation, etc. Applying such
methods to the above model, Kauffmann et al. (1999) generated a model galaxy
population that can be directly compared with observations. The high-luminosity
part of that population is shown in the top right panel of Fig. 6.5. Galaxies of
lower luminosity are shown in the bottom panels, divided according to their star
formation rate. Those with the low rate are shown in the left panel, and those
with the high rate are shown in the right panel. We can see that generally galaxies
follow the dark matter density distribution, but differences in the luminosity and
star formation rate lead to different spatial distributions of galaxies.
Thus far the models that have been compared with observations have been
mostly dark matter models. As we shall show in the later chapters (and as can
be seen from the above model), galaxies probably do not follow exactly the dark
matter distribution, so these comparisons cannot be considered entirely decisive.
However, with the rapid evolution of computer hardware and programming methods both semianalytical methods and direct gasdark matter simulations will soon
provide us with decent models of galaxy distribution.
256
=1
=0
85
128
10
198
2002 by CRC Press LLC
CHAPTER 7
P (Y (x1 ) y1 ; : : : ; Y (xn ) yn )
199
2002 by CRC Press LLC
188
for any number of points n. Because we will later work with Fourier transforms,
it is useful to think of the field Y as having complex values.
The properties of a random field are determined by its distribution functions.
For example, homogeneous random fields are defined by the condition that all distribution functions have to be invariant under a common shift of their arguments:
m(x) = E fY (x)g
(E denotes the expectation value). For a homogeneous random field m(x) does
not depend on x and must thus be a constant. This constant is usually taken to
be zero, but that is not always the case (e.g., in the case of asymmetric one-point
distribution functions).
The second moment is called the two-point noncentral covariance function
R2 (x1 ; x2 ) and is defined by
R (x ; x ) = E fY (x )Y ? (x )g :
2 1
The central two-point covariance function (x1 ; x2 ) is usually called simply the
covariance function and is defined by
?
(a star denotes complex conjugation). The covariance function is positive definite: for any collection of points fx1 ; : : : ; xk g and arbitrary complex numbers
fz1 ; : : : ; zk g the Hermitian form
XX
8
<X
E [Y (xi )
:
2 9
=
m(xi )] zi
;
0
is always real and non-negative. This condition limits the class of possible
200
2002 by CRC Press LLC
Random fields
189
correlation functions. It also allows us to represent the correlation function below as a Fourier integral.
For homogeneous fields the covariance function has to be the function of the
difference x1 x2 only. The value of the covariance function for zero lag (0) is
the variance of the field Y and is, evidently, real.
Yet another possibility to limit the class of random fields is to define isotropic
random fields by the requirement that the covariance function should not depend
on the direction:
The most popular random fields in cosmology are both homogeneous and isotropic,
and their covariance function (r) depends only on the distance between the two
points r = jrj.
The covariance function determines the (mean square) local properties of random fields. Namely, if the covariance function (r) of a homogeneous random
field is continuous at the origin (r = 0), then the field itself is continuous (in
mean square) everywhere. And if the 2k -th order partial derivative of the covariance function of a random field @ 2k (x)=@x21 : : : x2k exists and is finite at r = 0,
then the field itself has partial derivatives of the order of k .
The covariance function is the simplest case in a hierarchy of n-point covariance functions
(7.1)
(7.2)
In the event that we move one point (say, x3 ) far away from the other two, the
values of the field at that point become independent of those at the two other
points and the expectation value in (7.2) reduces to
190
2 (x1 ; x2 ) = 12 ;
3 (x1 ; x2 ; x3 ) = 123 ;
4 (x1 ; x2 ; x3 ; x4) = 12 34 + 13 24 + 14 23 + 1234 :
202
2002 by CRC Press LLC
(7.4)
Random fields
191
Y (x) =
IRN
e ixk dZ (k);
where the random field Z (k) is a field with orthogonal increments. The latter
means that different dZ (k) can be thought of as being independent. The field
Z (k) does not have to be differentiable; if it is, we can write its density as Ye (k),
and the above formula reduces to the usual (inverse) Fourier transform formula
Y (x) =
dN k
Ye (k)e ixk
:
(2)N
IRN
(7.5)
The field Y (x) and its Fourier representation Ye (k) form a Fourier transform pair,
and the Fourier representation can be found from the direct Fourier transform
Ye (k)
IRN
Y (x)eikx dN x:
There are many different Fourier transform conventions; those used here are common in cosmological papers.
The properties of the field Z (k) allow us to say that the Fourier amplitudes
Ye (k) are independently distributed, namely
n
(7.6)
N is the Dirac delta-function and P (k) is the spectral density of the field.
where D
The factor (2 )N is due to the adopted Fourier transform convention. It is easy to
remember, using dN k=(2 )N for the Fourier space volume element. If the field
Y (x) is real, Ye ? (k) = Ye ( k) and the above formula can be written as
n
192
(r) =
IRN
e irk dF (k);
where F (k) is the spectral distribution function of Y (r). It usually has a spectral
density P (k) (called power spectrum in astronomy) and the previous formula can
be written as
Z
dN k
(r) =
e irk P (k)
(7.7)
:
N
(2)N
IR
Because the correlation function is positive definite, the spectral density in (7.7)
has to be positive in its entire argument range (Adler 1981). This condition conforms to the definition of spectral density in Eq. (7.6).
For an isotropic random field the spectral density must be a function of the
norm k = jkj only. In this case the above formula can be simplified, integrating
over the angles first. For two dimensions, we get
(r) = 2
P (k)J0 (kr)
k dk
(2)2
(J0 (x) is the 0-th order Bessel function). For three dimensions,
(r) = 4
P (k)
sin(kr) k2 dk
:
kr (2)3
(r) = 2
P (k) cos(kr)
dk
:
2
In the Hilbert space formalism the covariance function and the power spectrum are the same function, only expressed in different bases. Application of this
approach can be seen in papers by Hamilton (e.g., Hamilton 2000).
The variance of an isotropic random field can be written as
2 = (0) = 4
P (k )
k2 dk
(2)3
(7.8)
and this has led to another frequently used form of the power spectrum in cosmology:
2 (k) =
1
P (k)k3 :
2 2
(7.9)
Using the above definition, we can write the total variance of the field (7.8) as
2 =
Random fields
193
thus, 2 (k ) gives the variance of the field per ln k . Because the power spectrum P (k) is a positive function of k, Eq. (7.7) describes all possible covariance
functions; any model covariance function should satisfy this relation. If we start
by choosing a model power spectrum, we achieve this automatically. However, in
cosmology we frequently use approximations for covariance functions that clearly
do not conform with (7.7). A typical example is the popular approximation for the
covariance function (r) = (r=r0 )
, which does not satisfy the above condition.
In this case we implicitly assume that our model is defined for a limited argument
interval.
As (7.6) shows, the power spectrum describes the second moment of a random
field. The third moment of the field can be written as
n
o
E Ye (k1 )Ye (k2 )Ye (k3 ) = B (k1 ; k2 ; k3 )(2)N DN (k1 + k2 + k3 ): (7.10)
The function B (k1 ; k2 ; k3 ) is called the bispectrum. Eq. (7.10) says that the third
moment is nonzero only for wavevectors that form closed triangles. This is the
result of homogeneity of the field. Because the correlation functions depend only
on the differences xi xj , one of the volume integrals in the Fourier transform in
(7.10) vanishes, except in the case k1 + k2 + k3 = 0 (see Peebles 1980, x 43). This
leads to the definition of the bispectrum as the Fourier transform of the irreducible
three-point correlation function:
B (k1 ; k2 ; k3 ) =
where the three wave vectors form a closed triangle, xij xi xj and we have
assumed that the field is homogeneous. The Fourier transform of the irreducible
four-point correlation function is called the trispectrum.
As we have seen above, many local properties of random fields depend on
the covariance functions and their derivatives. As the covariance function and
the power spectrum form a Fourier transform pair, these properties also can be
described by the spectral moments
Z
dN k
i1 i2 :::in =
ki1 ki2 : : : kin P (k)
(7.11)
:
N
(2)N
IR
=
IRN
P (k)
dN k
= R(0):
(2)N
In the case of real-valued random fields the covariance function is real valued
and the power spectrum is a symmetric function, P ( k) = P (k); hence, the
odd-order spectral moments are zero. The second-order spectral moments
Z
dN k
ij =
ki kj P (k)
3
(2)
IRN
205
2002 by CRC Press LLC
194
@ 2 R(x; y)
ij =
= R;ij (0)
@xi @yj x=y
(we have used the comma-notation for partial derivatives, e.g., @f=@xi f;i ).
Later on we shall need expressions for correlations between a Gaussian random
field and its derivatives, which can be found from the formula:
+
@
@ ++
+
Y (x) @
+ Y (x)
R(x) =
E
=
@ xi @ xj @ xl @ xm
@ xi @ xj @ xl @xm
x=0
N
P (k) d k : (7.12)
= (
ki kj kl
km
(2)N
IRN
This formula can be easily derived, using the spectral representations of the random field and of its covariance function. With appropriate choices for ; ;
;
we find that the field Y (x) and its derivatives Yi (x) are uncorrelated ( = 1; =
= = 0), and the second and first derivatives are also uncorrelated ( =
=
= 1; = 0).
1)+ i++ +
f (Y) =
(2)n=2
jSj
1
e[ 2 (Y
m)C
(Y m)T ] ;
mj )? g :
Cij = R(xi ; xj ):
As the covariance matrix of a multivariate Gaussian defines completely all the distribution functions (and, naturally, all the higher moments), the covariance function plays the same role for Gaussian random fields.
The same can be said about the power spectrum of a Gaussian random field.
The representation of a Gaussian random field Y (x) in the Fourier space Ye (k)
206
2002 by CRC Press LLC
195
Y (x; RF ) =
W (x x0 ; RF )Y (x0 ) d3 x0 ;
207
W(r)
W(k)
196
0
0
0.5
1.5
r/R
2.5
8
kR
10
12
14
Figure 7.1 The top-hat filter in the coordinate space (left panel) and in the Fourier space
(right panel). Compare the widths of the filters.
where RF is the filter scale the typical scale for a class of objects in which we
are interested. The Fourier image of the filtered field is then
f (k; RF )Y
e (k)
Ye (k; RF ) = W
WT H (r; RT H ) =
1
(R
(4=3)RT3 H T H
r);
where (x) is the theta-function. This filter smooths over a finite spherical volume
of radius RT H around a point, but its Fourier image is rather extended:
3
fT H (k ; RT H ) =
W
sin(kRT H ) kRT H cos(kRT H ) :
3
3
k RT H
(7.13)
We show this filter in Fig. 7.1 along with its Fourier image.
The second filter is the Gaussian filter that is equally smooth both in the coordinate space and in the wavenumber space:
1
exp( r2 =2RG2 );
(2)3 RG3
fG (k ; RG ) = exp( k 2 R2 =2):
W
G
WG (r; RG ) =
208
2002 by CRC Press LLC
197
W(r), W(k)
0.5
1.5
2.5
3.5
r/R, kR
Figure 7.2 The Gaussian filter, both in the coordinate space and in the Fourier space. The
y -axis units are different for the two cases.
VT H =
4 3
R :
3 TH
(7.14)
It is more difficult to define the effective volume for the Gaussian filter. A popular
convention is
p
VG = (2)3 RG3 :
(7.15)
These volumes are necessary when we wish to define objects by their masses, as
is customary in cosmology (during evolution the radii of objects change considerably, but their masses remain largely the same).
Although we all have seen Gaussian curves, we illustrate the Gaussian filter in
Fig. 7.2 so that you can compare it with the top-hat filter (see Fig. 7.1).
For a Gaussian filter the new field will be
Y (x; RF ) =
1
(2RF2 )3=2
02 2
e jx x j =2RF Y (x0 ) d3 x0 ;
giving a simple expression for the power spectrum of the filtered field:
2 2
P (k; RF ) = P (k)e k RF :
(7.16)
198
The basic form is that of the density (contrast) spectrum. We know from the
discussion of the dynamics of structure in Chapter 6 that growing modes of density, velocity, and gravitational potential are closely connected; thus, if we know
the spectrum of one of them, we can easily derive the others.
Contemporary cosmology predicts that the inhomogeneities that grow later into
the observed structure are generated during inflation. Cosmology predicts their
spectra, as well. There are many theoretical possibilities, but because we as yet
know little about the physics of inflation, most of these derivations have tried to
conform with previous intuitive ideas.
Intuitively, the best choice is the initial power-law spectrum P (k ) kn , as
there are no preferred scales for gravitation (the exponent n is usually called the
spectral index). In Chapter 6 we saw that in the comoving coordinates the density contrast and the fluctuating gravitational potential are connected via the
Poisson equation
( is the mean density of the universe). Thus, the Fourier components of these
fields are connected by
e = 4Ge(k )=k 2 ;
P (k) P (k)=k4 kn 4 :
This singles out the spectral index n = 1, because in this case there is equal
power in the fluctuations at any scale, meaning that the geometry of the universe
is equally wrinkled at any scale (the gravitational potential is the metric perturbation from the viewpoint of general relativity).
The n = 1 or P (k ) k spectrum is called the HarrisonZeldovich spectrum
and it is the most widely used form for the initial density contrast spectrum in
cosmology.
This is not the whole story, however. The initial fluctuations are influenced
by interactions between different species of particles in the early universe, by
interaction of matter and radiation, etc. Because the initial perturbations are small,
their dynamics are linear. It is easy to see that in the linear approximation the
Fourier components of the density contrast e(k ) evolve independently. Thus we
can write
e(k; t) =
D(t)
T (k; t; ti )e(k; ti );
D(ti )
where D(t) D1 (t) is the growing mode of the density perturbation we derived
in Chapter 6. The function T (k; t; ti ) defined by the above equation describes the
wavelength-dependent part of the evolution of the Fourier modes. It is called the
210
2002 by CRC Press LLC
199
transfer function and for a usual choice of the initial and current time moments it
does not depend on t and ti . The reason for that is that the physical processes that
depend on the wavelength of the mode work only for a limited time interval, from
the moment the mode enters the horizon until after recombination.
If we know the transfer function, it is easy to find the time-dependent spectrum
of the perturbations:
D(t) 2 2
P (k; t) =
T (k)P (k; ti ):
D(ti )
For an initial power spectrum P (k; ti ) = A(ti )k n the dependence on the initial moment is given by its amplitude A(ti ), which is usually determined from
observations (normalization of the spectrum). We shall describe that below.
Computation of the transfer function is rather difficult, but it is possible. It
involves integrating the Boltzmann equation describing various interactions, but
the perturbations are linear. The results of numerical studies are usually given by
approximation formulae. One of the most widely used forms is that for the cold
dark matter (CDM) universe, given by Bond and Efstathiou (1984):
n
i o 1=
(7.17)
200
2 (k) = H2
ck 3+n 2
T (k);
H0
(7.18)
where H is the rms amplitude of density fluctuations at the scale of the presentday horizon c=H0 . Because the normalization depends on the cosmological model,
they provide approximate expressions for H :
(
H =
1:94 10
1:95 10
5
0:785 0:05 ln
M e 0:95~n 0:169~n2 ;
M +
= 1;
M
5
0:35 0:19 ln
M e n~ 0:14~n2 ;
= 0;
M < 1;
M
(7.19)
e = n 1 is the tilt of the power spectrum, the difference of its initial
where n
slope from the HarrisonZeldovich value n = 1. The rms error they quote is 7%.
Another method of normalization is by the value of the (linearly extrapolated)
rms fluctuations of mass in spheres of certain radius. The radius adopted in cosmology for that is 8h 1 Mpc, which is considered to be approximately the smallest scale where the dynamics is still almost linear. This value is called 8 and is
given by
82 = 4
2
f 2 (k; RT H = 8h 1 Mpc)P (k ) k dk ;
W
TH
(2)3
0
(7.20)
fT H is the Fourier image of the top-hat filter (7.13). This value does not
where W
coincide with the observed rms variation of galaxy counts, which differs from 8
by the bias factor and by the effects of nonlinear dynamical evolution. Thus, 8
should be treated as only a normalization constant.
Of course, once we know the COBE normalization and the cosmological parameters, we can calculate 8 , so it is not really a free parameter. Using the 8
normalization means that we do not absolutely rely on known physical mechanisms of generating the initial power spectrum and allow for modifications that
can change its slope and amplitude near k = 2=8 h Mpc 1 . Thus, when determining the 8 normalization, the shape parameter in (7.17) is usually also
treated as a free parameter.
f 2 (k; R) drops sharply around k 1=R, the main
As the window function W
2
contribution to the (R) comes from the same region because of the volume
effect. Peacock (1999) gives a useful formula for quick estimation of 8 for CDMlike spectra:
82 2 (keff );
keff = 0:172 + 0:011 ln2 ( =0:34) h Mpc 1 :
The number density of clusters of galaxies is sensitive to the overall amplitude
of the power spectrum, so the 8 normalization is usually found on the basis of
cluster massnumber density distribution. There have been many attempts to do
212
2002 by CRC Press LLC
201
Figure 7.3 The normalization factor 8 for flat cosmological models. The hatched region
shows their result with rms error limits; lines show the results of different authors. (Reproduced, with permission, from Pierpaoli, Scott, and White 2001, Mon. Not. R. Astr. Soc.,
325, 7788. Blackwell Science Ltd.)
this and, in general, the results agree. In recent years 8 is mainly found on the
basis of the X-ray temperature distribution of rich clusters of galaxies (the Xray temperature is determined by the mass of a cluster). We refer the reader to a
careful analysis of methods used, their inherent uncertainties, and data problems
in Pierpaoli, Scott, and White (2001). They give detailed fits for 8 for different cosmologies and shapes of the power spectra. For a flat cosmological model
(
M +
= 1) and typical power spectra (n = 1 and = 0:23 in (7.17 above)
the result is
8 = 0:495
M0:60 :
For comparison, Viana and Liddle (1999) find
= 0:56
0:47 :
202
0.1
0
-1
P
-2
-2
-3
-3
-1
log( (k))
-3
log(P(k)Mpc )
1000
L (h-1Mpc)
100
10
-4
-4
-5
-5
-6
-6
0.001
0.01
0.1
10
-7
100
-1
k(h Mpc )
Figure 7.4 The density contrast power spectrum in a CDM cosmology. The lower x-axis
shows wavenumbers; the upper x-axis shows the corresponding spatial scales. Both the
spectral density P k and the spectral energy 2 k are in their real-life units.
()
()
The reason for that is simple in low-density models the growth of structure
stops at late cosmological epochs, so the observed clusters must have formed
earlier, and the power spectrum needs a larger amplitude.
Those two amplitudes (H and 8 ) restrict the overall behavior of the density
power spectrum rather well. For studies of galaxy clustering frequently only the
8 normalization is used, because it describes the power spectrum at typical clustering scales.
We show a typical density power spectrum in Fig. 7.4. It was calculated using
the coefficients obtained by Jenkins et al. (1998) by comparing the exact transfer
function with the form (7.17):
= 0:21, a = 6:4h 1 Mpc, b = 3:0h 1 Mpc,
c = 1:7h 1 Mpc, = 1:13. The normalization was determined by requiring
8 = 0:9.
We see that the spectral density P (k ) has a maximum at rather large scales
(around 200h 1 Mpc), and drops rapidly at small wavelengths. We also see that
the variance per logarithmic wavelength interval tends to a constant at small
scales; thus, the initial density distribution is highly irregular.
This is usually referred to as a hierarchical picture. The spherical collapse solution in Chapter 6 gives the collapse time as tcoll = 12tc = 12Rc =c, where Rc
is the characteristic radius of the sphere smaller objects collapse faster. Thus,
the density distribution can be visualized as a hierarchy of clouds of different size
and of different formation rate.
214
2002 by CRC Press LLC
203
Such a hierarchy of clouds is typical for the cold dark matter models. In once
popular hot dark matter models the transfer function decayed exponentially with
the wavenumber, starting at the characteristic scale of superclusters. In these models huge superclusters formed first and the statistics of objects was that of fragmentation, completely different from the present-day approach.
The spectra described above and shown in Fig. 7.4 define the density distribution at redshifts z 1000, at the start of the slow process of structure formation.
By the present day the power spectra have been distorted by dynamical processes,
but their overall form remains rather similar to the initial one.
Dynamical processes also can generate nonzero values of the bispectrum. In the
linear approximation the physical fields remain Gaussian, but in the second-order
approximation the bispectrum for density contrast is (Fry 1984)
Q123 =
B123
;
P1 P2 + P1 P3 + P2 P3
(7.21)
on the contrary, depends mainly only on the geometry of the triangle f123g. We
shall describe recent determinations of the bispectrum in Chapter 8.
7.4 Realizations of random fields
An instance of a random field is called a realization of that field. All the observed
fields are realizations of random fields. When building numerical models to compare with observations, we have to generate specific realizations of model fields.
Such realizations are also widely used as initial data for numerical simulations.
An overview of methods used for simulations of random variables, random fields,
and point processes in statistics is given in Ripley (1987).
The random fields used to model the fluctuating physical fields (density, velocity, etc.) in cosmology are usually Gaussian fields or local transforms of Gaussian
fields. As a realization of a random field is a set of field values Y fY (x1 ); : : : ;
Y (xn )g at n points, a realization of a Gaussian field can be considered as a multivariate Gaussian random variable of dimension n. Such a random variable is
completely specified by its mean m and covariance matrix C.
215
2002 by CRC Press LLC
204
Two methods to generate multivariate Gaussian variables are described in Ripley (1987). The most common method is to represent the covariance matrix as
C = SST . The square root S of the matrix C can be found by Cholesky decomposition (Press et al. 1992). Next we generate n independent Gaussian random variables Z fZ1 ; : : : ; Zn g and calculate the sum
Y = m + SZ:
It is easy to see that the multivariate random variable Y is the realization we seek.
This method, however, is suitable only for rather small n, because the decomposition of large matrices is computationally expensive.
Other methods for generating Gaussian random fields, listed in Ripley (1987),
use the central limit theorem. One example is Matherons turning band method,
which is used to generate realizations of homogeneous and isotropic Gaussian
random fields. This method starts by simulating a stationary one-dimensional
Gaussian process Y (x) with a known covariance function 1 (r). Then we select N -dimensional random rotations O and carry the values of the process on the
line over to the points in N -dimensional space covered by the rotated image of
the line, Y (Ox) = Y1 (x). The generated process is isotropic with the covariance
function N (r) (Ripley 1987):
Z
1
2 (N=2)
N (r) = p
1 (vr)(1 v2 )(N 3)=2 dv:
[(N 1)=2] 0
The relation between the covariance functions is especially simple for N
1 (r) =
= 3:
d
[r (r)]:
dr 3
205
x(j) = x
l=3
X
l=1
jl el ;
and el are the unit vectors in coordinate directions. Because of periodicity the
function Y can be represented as a discrete Fourier transform (see, e.g., Press et
al. 1992)
Y (j) =
1 X ix(j)k(l)
e
ye(k(l));
(2)3 l
(7.22)
where the Fourier amplitudes ye(k) are also defined on a lattice l = (l1 ; l2 ; l3 ); ln =
N=2; : : : ; N=2,
nX
=3
k(l) = k
ln un :
n=1
Here un are the unit vectors in the Fourier space, and the k -space lattice spacing
is k = 2=L. The finite number of lattice points in one direction limits the
range of the Fourier sum by the Nyquist frequency kN = N k=2. Comparing
Eqs. (7.5 and 7.22) we see that
206
( ) = k2 (for
( )=
variance
2 =
(k)3
P (k):
2
As our field is real, its Fourier amplitudes have to satisfy the relation
(7.23)
with the exponent n = 1 and the smoothing radius R = 0:3, on a 322 grid.
It is clear that different spectra lead to different types of random fields. This
is illustrated in Fig. 7.6, which compares realizations with pure power spectra,
P (k) = kn , but with the exponents n = 2 and n = 4. The larger exponent gives
more weight to fluctuations (Fourier components are essentially plane waves) with
higher wavenumbers jkj and with shorter wavelengths, making the realization
more erratic (although both realizations are smooth). The smaller exponent, in
contrast, weighs larger wavelengths more heavily, and leads to a rather smooth
realization even without a high-wavenumber cutoff.
218
2002 by CRC Press LLC
207
2x 2 2
f (x) = 2 e x = ;
(7.24)
where
2 = (k)3 P (k):
The integral probability function for (7.24) is
(7.25)
thus the function exp( jyej2 = 2 ) has a uniform distribution on the unit interval.
This gives another means for generating the Fourier amplitudes ye: generate a random value from the uniform distribution and transform it to the module, using
(7.25). Then select a random phase and use that to generate the real and imaginary parts of the complex amplitude by
208
(2)3
If we convolve the white noise field with the function G(x), the resulting field
Z
Y (x) = G(x x0 )YW (x0 ) d3 x0
IRN
where " > 0. As the variance of the field is given by a similar integral,
Z
dN k
;
2 =
P (k)
N
(2)N
IR
we see that the finiteness of the variance does not yet guarantee the smoothness
of realizations. For an isotropic 3-D Gaussian random field the above condition
limits the high-wavenumber behavior of the spectral density, demanding that
k!1
220
2002 by CRC Press LLC
Non-Gaussian fields
209
1
R(x1 ; x2 ) = (jx1 j + jx2 j
2
jx1
x2 j) :
The realizations of such fields are more difficult to construct, although they are
widespread; two-dimensional index- fields serve as the basis for fractal landscapes. These are usually built by fractal interpolation (the random midpoint displacement method). In this method the values of the field are generated initially
at a coarse grid (the four corners of a square already represent a grid) and these
are interpolated to successively finer grids, adding to the interpolated values random Gaussian displacements with the variance chosen to satisfy the correlation
function. The construction of fractal landscapes is well described in the literature
and there are many freeware programs to generate them. The best-known collection currently can be found at the Spanky Web site (Spanky). We also refer the
interested reader to a volume edited by Peitgen and Saupe (1988). Fig. 7.7 shows
an example of such a landscape, a realization of the two-dimensional isotropic
Brownian motion.
As Fig. 7.7 shows, this surface is really extremely erratic. In fact, this is not a
normal surface any more, because realizations of index- fields are fractal with
the fractal dimension D = 3 (for two-dimensional fields). The dimension of
the surface in Fig. 7.7 is D = 2:5, and this surface is halfway to filling the volume
it occupies (Mollerach et al. 1999).
Thus, Gaussian random fields can be used to model fractal distributions of matter in the universe, should the need to do that arise.
7.5 Non-Gaussian fields
General non-Gaussian random fields are difficult to characterize. The fields that
are used in cosmology are all derived from a Gaussian field, with a long list
given by Coles and Barrow (1987). An example is a lognormal field L(x) =
exp(G(x)), where G(x) is a Gaussian random field. This field has been described
in detail by Coles and Jones (1991). Although the field L(x) is homogeneous, if
we start from a homogeneous Gaussian field, it is not wise to consider L as a
zero-mean field, as its one-point distributions can be very asymmetric. There is a
simple relation between the covariance functions:
210
=25
where RL (r) is the covariance function of the lognormal field, RG (r) is the covariance function for the Gaussian field, and we have supposed that the Gaussian
(hence also the lognormal) field is isotropic.
Another rather well-studied class of fields are the 2 fields, also described by
Adler (1981). A parameter-n 2 field Yn (x) is defined as
n
X
Yn (x) = G2i (x);
i=1
where Rn (r) is the covariance function of the parameter-n 2 field and RG (r) is
the covariance function of the component Gaussian fields.
The Rayleigh fields are similar to the 2 fields and are defined as
RRn (x) =
Yn (x):
Several versions of these fields for specific n are described by Coles and Barrow
(1987). Sheth (1995) gives the one-point distributions and correlation functions
of these fields for the general case.
222
2002 by CRC Press LLC
211
hG2 (x)i ;
Y (x) = Y (0) +
1X
Y; x x ;
2 ij ij i j
223
2002 by CRC Press LLC
212
where i; j = 1; : : : 3 label the coordinates and Y;ij = @ 2 Y=@xi @xj (0) are the
second derivatives of Y at the maximum. The first derivatives Yi are then
Y;i (x) =
Y;ij xj :
(7.26)
dY d3 x
np (Y ) =
(7.27)
where the integration is over the region M in the space of second derivatives
that corresponds to the maxima of Y , namely, where the eigenvalues of Y;ij are
negative. (Note that our notation for peak number densities differs from that of
BBKS, who use N for differential and n for integrated peak number density.)
The joint probability density of the Gaussian field and its derivatives is also
Gaussian:
1
p
e
(2)N=2 jCj
yC
1 yT
where the vector y stands for the collection of arguments Y; Y;i ; Y;ij . Its dimension is 10, as there are only 6 independent second derivatives, and the matrix C
is the 10 10 correlation matrix of y. The components of the correlation matrix,
the correlations between the field and its derivatives, are given by the correlation
function and its derivatives or, alternatively, by the spectral moments, as we described above. For a real field the odd-ordered spectral moments are zero, which
simplifies the matrix.
The correlation matrix can be found using the general formula (7.12). Using
these amplitudes and the mapping y1 = Y; yi+1 = Y;i (i = 1; 2; 3); yi+4 =
Y;ii ; y8 = Y;23 ; y9 = Y;13 ; y10 = Y;12 , the nonzero components of the correlation matrix can be written as
C00 = 02 ;
1 2
C0i =
; i = 5; 6; 7;
3 1
1
Cii = 12 ; i = 2; 3; 4;
3
224
2002 by CRC Press LLC
(7.28)
213
1 2
; i = 5; 6; 7;
5 2
1 2
; i = 8; 9; 10;
Cii =
15 2
1 2
Cij =
; i; j = 5; 6; 7; i 6= j;
15 2
Cii =
where i are the spectral amplitudes. These are defined in cosmological literature
as
Z
1 1 2i
k P (k) dk:
i2 = 2
(7.29)
2
These are usually called spectral moments and they differ from the moments
used in the theory of random fields (see 7.11).
The next quantity we have to know to calculate the integral (7.27) is the volume
element d6 Y;fij g . Bardeen et al. (1986) show that this can be written as
np ( ) =
2
1
e =2 G(
;
):
(2)2 R?3
(7.31)
=
the function G is defined as
G(
; x? ) =
and
F (x; ; x? ) =
Y
;
0
F (x;
; x? ) dx;
(7.32)
2
2
1
e (x x? ) =2(1
) f (x):
2
2(1
)
(7.33)
The parameters
and R? characterize the shape of the spectrum and are defined
as
12
;
0 2
p
R? = 3 1 :
0
=
225
2002 by CRC Press LLC
(7.34)
(7.35)
214
R? =
and
2 =
6
R ;
n+5 F
n+3
:
n+5
f (x) =
(x3
r
3x)
2
+
5
"
erf
5
x + erf
2
5x
22
2
31x2 8
x2
+ e 5x =8 +
4
5
2
!#
+
2
8
e 5x =2 :
5
x=
r2 Y=2 :
(7.36)
As shown by Mann, Heavens, and Peacock (1993), the formula (7.31) can be
integrated analytically to get the total peak number density over a level
Np ( ) =
where I ( ) is an integral
I ( ) =
f (x)e
I ( );
82 R?3
x2 =2 erfc
(7.37)
!
x
p
dx:
2(1
2 )
(7.38)
The peak number density np ( ) and the total density Np ( ) (the total number
density of peaks with height ) are shown in Fig. 7.8 for various values of the
concentration parameter of the spectrum
. We see that the amplitudes of peaks
can reach up to 5 , and that there can be peaks with height much lower than the
mean density of matter, down to 3:5 . There are no low-density peaks only in
the limiting case of a very sharp spectrum,
= 1.
226
2002 by CRC Press LLC
215
0.012
0.018
0.016
0.01
0.014
Np()R
0.012
np()R
0.008
0.006
=1
0.004
=0
=1
0.006
0.004
=0
0.002
0.01
0.008
0.002
0
-4
-3
-2
-1
-4
-3
-2
-1
Figure 7.8 Peak number densities for Gaussian fields with a different spectral parameter
.
The left panel shows differential number densities dependent on the normalized peak height
. The different curves correspond to different
, ranging between the limiting values
and
with the step 0.2. The right panel shows the total number density (for all peaks
with height larger than ) for the same choice of
.
=0
=1
We see also that the mean peak density grows with the concentration of the
spectrum, ranging from 0 for
= 0 up to about 2.2 for
= 1. The right panel
of Fig. 7.8 shows us that the total peak number density does not depend on the
concentration, but only on the correlation length R? . This also can be seen from
the formula (7.32) the integral of exp( 2 =2)G(
;
) over all peak heights
does not depend on
.
The total peak number density can be found exactly:
29 6 6
Np ( 1) = 2 p
8 125R?3
0:00162R? 3 ;
one peak per about (9R? )3 volume. If we suppose that the object connected with
the peak has a volume of the sphere of a radius of R? , the peaks occupy roughly
1=50 of the total volume for any spectra.
In the high peak limit, the formulae for the number densities simplify:
2
3
( 3 3 )e =2 ; ! 1;
2
3
4 R?
2
3
( 2 1)e =2 ; ! 1:
Np ( ) =
2
3
4 R?
np ( ) =
227
2002 by CRC Press LLC
(7.39)
(7.40)
216
Y (x) = Y0
i x2i ;
Y0 = Y (0);
if we orient the coordinate system along the principal axes (the eigenvectors of
Y;ij ). At a peak the eigenvalues i are all positive and they define the semiaxes of
the profile:
2(Y0 Yc ) 1=2
ai =
;
i
where Yc is a constant value of the field.
The steepness of the profile (its curvature at the center) is
r2 = i i = I1 ;
where I1 is the first invariant of the matrix Y;ij . Let us order the eigenvalues by
1 2 3 . The asymmetry of the profile is described then by its ellipticity
(in the x1 ; x3 plane)
e= 1 3
2I1
22 + 3
:
p= 1
2I1
The last parameter has values p 2 ( e; e); for prolate spheroids p = e and for
oblate spheroids p = e.
Bardeen et al. (1986) derive the joint probability distributions of the height of
the peak , its steepness I1 , and the asymmetry parameters e and p. The formulae
are rather cumbersome and we will not repeat them here. The main result is that
the most probable value of p is close to zero (2 (1 + 3 )=2), but the ellipticity
e is definitely larger than zero; the peaks are triaxial. The higher the peak, the more
spherical it is the most probable ellipticity is
emax
p1 :
(7.41)
5
Probably more important than the initial density profile is its evolution during
the collapse. In the Zeldovich approximation this evolution is determined by the
velocity potential (x). Let us diagonalize the (symmetric) deformation tensor
;ij and denote these eigenvalues again i (they are different from those that
describe the density profile) and order them. Doroshkevich (1970) found a rather
simple expression for the joint distribution of the eigenvalues:
153
p exp
p(1 ; 2 ; 3 ) =
8 506
(1 2 )(1
228
2002 by CRC Press LLC
3I12 15I2
+
02 202
3 )(2 3 );
217
where I1 and I2 1 2 1 3 2 3 are the first and the second invariants of the
2 3 )
deformation tensor. This formula shows that spherical collapse (1
is rather improbable.
It is easy to see that the first invariant of the deformation tensor is the density
contrast . Defining the ellipticity and prolateness of the deformation tensor as we
did above for the density profile, an exact result for the joint conditional probability density of e and p given can be found (Sheth, Mo, and Tormen 2001):
f (e; pj ) =
(
3 10
5 2 (3e2 + p2 )
2
close to that for the density asymmetry distribution above (7.41). This formula is
valid for a general point in the flow; adding the peak constraints would make it
more complex, so it serves here as an illustration.
7.6.3 Clustering of peaks
The motivation for developing the peak statistics is that we can hope to identify
the peaks with observed objects and thus to infer the properties of the random
field itself. Apart from the number densities of peaks, the next simplest observational statistics is the correlation of the peaks. To model that we have to find the
covariance function of the peak density field.
In order to find the covariance function of maxima we have to calculate the
joint probability density of having maxima at two points separated by r. If the
almost diagonal
one-point density studied in the previous section had a
correlation matrix (7.28), the correlation matrix for the present distribution is, nat matrix. This matrix, however, has many nonzero, nondiagonal
urally, a
components, because correlations between the field and its derivatives at different
points do not vanish. Bardeen et al. (1986) give recipes for a numerical solution of
this problem. We shall use the approximation proposed by Kaiser (1984), calculating the density correlations of regions where the field is higher than some fixed
level. When that level is small, these regions do not tell much about the maxima
of the field, but when the level rises, it delineates the maxima better and better.
If we suppose that maxima populate overdense regions roughly with a uniform
number density, we can approximate the correlation function of maxima by that
of the density correlation of overdense regions.
The overdensity correlation function is much easier to calculate because that
depends only on the values of the density field at two points and does not involve
the first and second derivatives.
10 10
20 20
229
2002 by CRC Press LLC
218
In order to develop the joint distribution of the density contrast field (a homoge; at two points,
neous and isotropic zero-mean Gaussian field) i xi ; i
we shall follow Peebles approach (1993) and introduce the auxiliary variables
+ 1 2 and
1 2 . These variables are also Gaussian and they are
independent, as
= ( ) =1 2
= +
g = E f12 g E f22 g = 0:
E f+
Their variance can be found as
()
p(+ ;
) = 2p2 (0)1
= 2[ (0)
"
2 (r)
+2
( )]
#
(r)
Substituting +; by their definitions via 1;2 and taking into account that the
determinant of the Jacobian of the transformation 1;2 ! +; equals one, we
find the joint probability density of 1;2 :
1
p(1 ; 2 ) = p 2
2 (0)
2 (r)
exp
"
= (0)
P2 ( ) =
1Z 1
p(1 ; 2 ) d1 d2 :
(7.42)
(; )
1 + (r; ) = P2 (r; )=P12 ( );
()
where P1 is the one-point probability for the density to exceed the level . For
an exact procedure to find the integral in (7.42) and the exact correlation function
we refer readers to the paper by Jensen and Szalay (1986). They represent the
function as an infinite series that has to be summed numerically. We shall derive
below an asymptotic expression, following Padmanabhan (1993).
230
2002 by CRC Press LLC
219
For high peaks ( ) the integral in (7.42) simplifies. After a change of variables yi i = the integral (7.42) can be written as
"
1Z 1
exp
y22 =2
exp
"
exp
y22 exp
"
y12 + y22
(r)y2 2
2
y1
1 we get
(r)y2 2
2
which, after discarding quadratic terms in , gives a similar integral for y2 . Integrating that, we get
2
2
e e (r) ;
P
2(
) = 21
P1 ( ) =
we get, finally, the correlation function
p1 e
2
2 =2 ;
(r )
(r; ) = exp 2
(0)
1:
(7.43)
(r; ) = 2
(r)
(0)
= 2 (r):
(7.44)
This expression was obtained first by Kaiser (1984). Although the formulae (7.43,
7.44) are approximate, the comparison made by Jensen and Szalay (1986) shows
that they do not differ much from the exact results for the correlation of highdensity regions.
Another approximation made above was to substitute the full density field for
the peak density field. Nevertheless, formula (7.44) has the same form as the
asymptotic formula for the correlation of peaks, obtained by the exact procedure.
231
2002 by CRC Press LLC
220
e =
by an effective
x
;
2
(7.45)
(r ; ) =
hei2 (r):
(7.46)
2
Here hei is the mean value of the effective density level (see (7.31, 7.32)) over the
peak number density distribution,
0
x
d 0 dx;
2
0
(1
)
F
(
x
;
;
)
0
? p
where F (x) is given by (7.33) above. This can be integrated over 0 to get
p
Z 1
x)
2
x2 =2 exp
f
(
x
)
e
hei = p
2(1
2 ) dx;
(1
2 )I ( ) 0
2
e =2
0
()
()
()
()
p (x) =
b (x)=0 :
np (p ; x) = np
b (x)=
232
2002 by CRC Press LLC
p ( ) b (x)
np ( ) 1 n 1( ) dnd
:
p
221
This means that we can introduce the peak overdensity field, similar to the
background overdensity, by
2 g
;
L np = bL (p ; 0 )b = p
(7.49)
p b
where we have introduced the peak overdensity bias bL (p ). The superscript L
means that the above formula describes the situation in the initial (Lagrangian)
space (we shall change over to real space below).
As we see, the local peak overdensity is enhanced by the factor bL p ; 0 compared to the background overdensity b . For high peaks, p2 g , this translates to
the factor p2 =02 for the correlation function, reproducing the amplification found
above.
To find the peak overdensity in real (Eulerian) space we have to account for the
movement of the peaks. We can do that by considering the peaks as markers in
the background field. In this case their density grows together with the density of
the background,
nEp nLp b ;
where the superscript E denotes Eulerian quantities. For overdensities,
nEp
nLp
b
nLp b :
= (1 +
)(1 + ) 1 =
+
We have assumed that the background overdensity b 1; because it describes
large scales, this assumption is justified. We have also assumed that its Eulerian
value coincides with the Lagrangian one. This is a good approximation; an exact
treatment can be found in Catelan et al. (1998). The peak overdensity in real space
is now
!
E np = bE (p ; 0 )b =
1 + p
b ;
(7.50)
222
As correlation functions scale as the square of the density contrast, we get the
peak correlation function
2
p (r; ) = bE (p ; 0 ) (r):
(7.51)
This formula can be used for all peak heights, if we calculate the exact g (7.48).
The behavior of the peak bias can be analyzed, defining a typical peak amplitude ? by
?2 =2 = g:
The expression for bias (7.50) tells us that the peak height ? separates the different biasing regimes. Higher peaks are biased (their overdensity is higher than the
background, mass density), but lower peaks are anti-biased (their overdensity is
smaller than that of the background).
The discussion above is not rigorous, because it is difficult to separate the background and peak fields. A natural way would be to generate the background field
by filtering the initial field and to get the peak field as the difference between the
initial field and the background. In that case, however, the two fields will not be
statistically independent. A perfect case would be if the power spectrum of the
initial field had a wide gap, a region of wavenumbers where it is zero. The spectra that we have seen in cosmology, however, are not even remotely like that. An
additional requirement is that there should be enough large-scale power to form
a background. For typical cosmological density power spectra this requirement
is true for a rather wide wavenumber interval, from high wavenumbers down to
about k : hMpc 1 (L h 1 Mpc) (see Fig. 7.4).
Thus, the peakbackground split remains a heuristic explanation. But it is useful to know that the higher the large-wavelength plateaus of a random field, the
higher the peak number density in these regions.
01
60
223
larger than their possible total displacement, so galaxy clusters retain their identity.
The next problem is to find the connection between the basic arguments of
the peak theory, the radius of the smoothing filter RF and the height of a peak
=0 ( denoting here the smoothed density contrast field) with the observed
parameters of galaxy clusters. There are several methods for that; we shall follow
the approach of Cen (1998).
As the initial perturbations are small, the mass of a cluster is determined mainly
by the smoothing radius of the density field. The relationship between these quantities can be derived, comparing the mass of an object inside the filter profile
MF
3=2 RF3 M c
(7.52)
= (2 )
(c is the critical cosmological average density) with a mass of a virialized spherical cluster
Mvir =
4 R3 vir ;
3 vir
178
(7.53)
(7.54)
(see Chapter 6). In this picture all peaks of a field smoothed by a filter of radius
are assigned the same masses. As the mass range of real galaxy clusters is
wide, we have to describe the cluster population by a set of smoothed fields for
different filter radii.
The filter radius for a particular value of the virial mass can be selected, obviMvir , which leads to the relation between RF and
ously, by demanding MF
rvir :
RF
RF3
2 0:7 3
= 178
3
M Rvir :
(7.55)
The masses of Abell clusters MA are defined as the mass inside the Abell cluster
: h 1 Mpc. This is a fixed radius at the present epoch; the filter
radius RA
radius (7.55) derived above describes the region in the initial density field that
has collapsed to form the cluster. The larger this radius, the larger the mass of
the galaxy cluster. To relate the integrated mass of an Abell cluster to the virial
mass (7.53) we have to assume the cluster density profile. Cen (1998) chose a
power-law density profile r r , which gives
RA 3
MA Mvir
(7.56)
:
Rvir
= 15
()
Substituting here (7.53) we find the relation between Rvir and MA and using
(7.55), we finally get a relation between RF and MA . This expression is rather
cumbersome; we refer the reader to the paper by Cen (1998). Apart from the parameters of the cosmological model, the only adjustable parameter in this relation
235
2002 by CRC Press LLC
224
23
is , which Cen found to be : , comparing the cluster mass functions from numerical simulations with the peak number density relations.
The minimum peak height for a peak to be virialized by the present time and to
: for spherical collapse (Chapter 6). Cen
be counted as a galaxy cluster is c
argues that as real collapse could differ from that (generic collapse is anisotropic),
it should also be treated as a free parameter. He finds by comparing the simulation cluster mass functions with peak theory results that it also has practically a
: for all cosmological models.
constant value c
If the two parameters are fixed, there is no more additional freedom in the
theory and it can be applied directly to observations. This is basically the way
filter radii are connected with cluster masses in all other studies, including those
based on the PressSchechter formalism that we describe below.
Analyzing the behavior of the cluster mass function in the peak theory, Cen
found that it depends mainly on the normalization of the spectrum 8 , on the
density parameter M , on the shape of the power spectrum, and slightly on .
The peak theory normalization
= 1 68
= 1 40
8 0:5
0:43 ;
is slightly lower than that obtained from the application of the PressSchechter
theory (see Section 7.3).
It is much easier to restrict the range of cosmological models that are allowed
by observations using the analytical peak theory than by extensive and costly
numerical simulations. Cen (1998) carried out the first analysis of the parameter
space and found six CDM models that possibly bracket all observational results.
He also compared the correlation function of simulated clusters with the peak
theory formula
2
h
ei2
+ 1 (r);
p
2
which is the formula (7.46) derived above, modified for gravitational motions in
Eulerian space, and found a good agreement. A general agreement of the amplification of correlations predicted by the peak theory was found earlier.
Another prediction of the peak theory is the dependence of the correlation function and the peak number densities on the density level that defines the peaks.
Peaks with larger have higher amplitudes, they collapse earlier and turn into
more massive (richer) galaxy clusters, because there is more time to accumulate
galaxies by accretion. The richer the clusters, the smaller their number density
Np 1=3 ).
(this can be described by the mean distance between clusters, dc
So we can compare the predicted amplification versus density level relation with
observations. As the amplitude of the correlation function is described by the correlation length r0 , this test is presented as r0 versus dc dependence for different
cluster samples.
236
2002 by CRC Press LLC
225
Figure 7.9 The correlation radius mean cluster distance dependence for cluster samples
of different richness. The line shows the linear relation r0 = 0:4dc of Bahcall and West
(1992). The data do not confirm this trend. (Reproduced, with permission, from Croft et al.
1997, Mon. Not. R. Astr. Soc., 291, 305313. Blackwell Science Ltd.)
This test has an interesting history. In 1986, Bahcall and Burgett found on the
basis of a few (four) samples a linear relation
r0 0:5dc
(7.57)
between the cluster correlation function and the mean distance between the clusters in a sample. Attempts to reproduce this relation on the basis of the peak
theory failed. As shown by Mann, Heavens, and Peacock (1993) and by Croft
and Efstathiou (1994), the peak theory gives substantially smaller amplification
factors (smaller r0 for a given dc ) than the relation quoted above. This can be
easily seen for a power-law galaxy correlation function r r=r0 2 the
cluster correlation length will be amplified by = , but the mean cluster distance
2 = , so the dc versus r0 relation should level off with
by np 1=3 1
the growth of .
This contradiction was thought first to be a deficiency of the peak theory. However, attempts to explain the relation (7.57), using detailed numerical models of
the formation of the clusters (Croft and Efstathiou 1994 used both methods), led,
basically, to the same result the r0 versus dc dependence leveled off for larger
dc . If the dependence (7.57) was right, the sites of formation of rich clusters
should be more correlated than possible in a Gaussian field, and non-Gaussian
distributions had to be invoked.
() (
( )
exp( 6)
237
2002 by CRC Press LLC
226
PressSchechter method
227
Then the peak patches that are totally inside larger patches are eliminated, and
the mass divided between partially overlapping peaks (the final solution of the
cloud-in-cloud problem). The Lagrangian space is divided in this way into
nonoverlapping patches of different mass.
2. The Lagrangian patches are moved into their final positions, using either the
Zeldovich or higher-order approximations and the initial velocity fields at the
patch, and patches are merged, if necessary.
3. The final density profiles for patches are calculated, based on the ellipsoidal
collapse of a patch. The ellipsoidal collapse is solved here and above for selfconsistent external tidal fields.
The procedure can be extended to include large-scale background power and
to form objects at different redshifts, to simulate observations along a light cone.
Although it is rather complex and not easy to program, it runs much faster than
the usual N -body programs and can simulate larger volumes. The peak patches
are groups of objects (galaxies). The comparison of the method with a N -body
simulation, using identical initial data (Bond and Myers 1996b), shows that it
reproduces the simulated groups very well. This could mean that in the future,
when we return from the study of large volumes to that of individual objects, we
have to use the language of peak patches; peak statistics is waiting for its time.
7.7 PressSchechter method
The PressSchechter method was developed to predict the mass distribution of
hierarchically formed objects on the basis of initial power spectrum (Press and
Schechter 1974). At first glance, this does not seem to be a problem of spatial
statistics. However, this method is based on a specific picture of the spatial distribution of galaxies, has been used to construct statistical models for that distribution, and has been extended to predict correlations of objects of different mass. It
is also one of the most used tools in cosmology.
The PressSchechter approach starts searching for objects of mass M by
smoothing the initial overdensity field by a filter of radius R R M . This filter
VG R
is usually taken as Gaussian, in which case, conventionally, M R
3=2 R3 . Another assumption is that regions of overdensity C have collapsed into observable objects. Such an overdensity limit follows from the spherical collapse solution described in Chapter 6. In order to describe mass distributions we have to consider fields with all smoothing radii at a given moment (at
present) and assume that all objects have the overdensity C . The overdensity
field is usually normalized to its value at the present epoch (the value it would
have if the evolution were linear). The typical numerical value of the collapse
: .
limit is then the well-known C
= ( )
( )= ( )=
(2 )
= 1 686
239
2002 by CRC Press LLC
228
The usual assumption of Gaussianity of the initial overdensity field gives the
probability that a given point belongs to a collapsed object (a massive halo)
PG ( C jR) =
( )
1 1
2
erf
p C
2(R)
(7.58)
where 2 R is the variance of the filtered field. A concern is that the maximum
probability given by this formula is 1/2 and not unity (although it is not clear at
all why all the mass in the universe should be concentrated in objects). Press and
Schechter introduced a fudge factor and wrote for the fraction of mass in halos
with mass greater than M (the integral mass distribution)
F (M ) = 1
(= 2);
where = C = (M ). The mass function n(M ) is defined as a comoving number
density of halos of mass M , so
erf
dF
;
dM
Mn(M ) =
n(M ) =
d
M2 d
2
ln (M )
ln M
2 =2
e
:
(M ), we get
(7.59)
= 3
(8 )
2 (M ) = 4
(; )
f (k ; M )j2 P (k ) k dk ;
jW
(2)3
(; )
f
where W
D (0)
C (t) = 1 C :
D1 (t)
Early comparison with numerical experiments showed that the formula (7.59)
describes simulations well, especially if we treated the critical overdensity level
C as a free parameter. The main problem then was the explanation of the fudge
factor. The problem was resolved after Peacock and Heavens (1990) and Bond
et al. (1991) showed that this factor could be derived, if one considered a fourdimensional overdensity field x; R , where R is the filtering radius. For a set of
widely spaced filtering radii the filtered fields could be considered essentially independent Gaussian random fields. For a fixed x the R-dependence of the amplitudes can be used to determine the mass associated with that point. The procedure
is to start from the maximum R possible and to assign to each point x the filtering radius R when the smoothed overdensity at that point (the trajectory R x )
( ; )
240
2002 by CRC Press LLC
PressSchechter method
229
first upcrosses the collapse level C . This filtering radius gives the largest mass
around that point that has collapsed by the present time, hopefully also destroying any substructure during collapse. This eliminates the possibility of multiple
counts of a mass point (the cloud-in-cloud problem).
The probability for a given point to belong to a halo of filtering radius R or
larger can be written as (Peacock and Heavens 1990):
P (R) = PG ( C jR) =
Z C
dPG
Pup (; C ) d;
1 d
(7.60)
( )
where PG is given by (7.58) and Pup is the probability that R could exceed
the threshold C at some filtering radius larger than R (the upcrossing probability). Thus, the statistics of the excursion sets of the trajectories of the smoothed
overdensity R x can be used to derive the mass functions.
The nature of these trajectories depends on the filters used. Peacock and Heavens (1990) derived approximate expressions for P R for different filters. In all
cases the result differs considerably from the simple PressSchechter assumption,
usually predicting more small-mass halos. Bond et al. (1991) demonstrated that
the PressSchechter formula, together with the fudge factor 2 can be obtained
only in the case of a sharp k -space filter. In this case the trajectories execute Brownian walks the increments of are formed by a sum of independent Fourier
amplitudes and are independent. They recovered the PressSchechter mass function with an ideal normalization (unity) by calculating the fraction of trajectories
that have been absorbed by an overdensity barrier C by the time 2 R .
The independence of the overdensity increments in this picture shows that it is
the mathematical expression of the peakbackground split approximation. While
for realistic filters the filtered density fields for different filter widths are correlated, the sharp k -space filter eliminates this correlation by definition. Thus, in
addition to other assumptions, the PressSchechter picture also includes the assumption of the exact peakbackground split.
Bond et al. (1991) also derived the merger probability for masses, collapsed
and virialized by an earlier moment t1 , to be included now in a larger mass. It is
described by a similar Brownian walk with two absorbing barriers. The fraction
of the mass of a large region of mass M0 with variance 02 and overdensity limit
0 that has been brought by smaller halos of mass M1 with the present linearly
extrapolated overdensity limit 1 and variance 1 , is
( ; )
( )
( )
f (1 ; 1 j0 ; 0 )
d12
dM =
dM1 1
2 d 2
1 dM :
p1 (21 2)03=2 exp 2((1 2 0)2 ) dM
1
2 1 0
1
0
1
Mo and White (1996) used that relation to predict the halo bias and the enhancement of spatial correlations in the PressSchechter picture. The average number
241
2002 by CRC Press LLC
230
M0
d2
f (1 ; 1 j0 ; 0 ) 1 dM1 :
M1
dM1
This gives the overdensity of M1 halos in M0 :
n (1j0)
1;
hL (1j0) = h
n(M1 )V0
where n(M ) is the standard PressSchechter mass function by (7.59) for the critical level C (t1 ) and V0 is the volume occupied by the mass M0 (a sphere of
radius R0 ). For a given power spectrum and filter the variance 2 (M ) and the
nh (1j0)dM1 =
12 ) and j0 j 1 is
2 1
;
hL = bLh 0 = 1
1 0
2 1
bEh = 1 + 1
:
1
(7.61)
We see that it does not depend on the characteristics of the larger region.
As these bias factors are mean values over regions, the PressSchechter approach can predict only average correlations, defined as
Z R0
(R0 ) = 3
(y)y2 dy:
R0 0
2
h (r) = bLh (M1 ; 1 ) (r):
For scales sufficiently larger than R1 the same relation applies to the correlation
functions themselves.
In order to define the PressSchechter biasing and anti-biasing regimes, we
shall define a mass M? by 1 M? . Using this approach, we vary masses of
halos. For larger masses the variance 2 M is smaller, so by (7.61) more massive
halos than M? are (positively) biased and smaller halos are anti-biased.
Comparing the bias factors from the PressSchechter approach with those obtained in the peakbackground split approximation of the peak theory (7.49, 7.50)
we see that they are very similar, the only difference being in the constant g that
is unity for the PressSchechter picture and can reach the value 3 in the high-peak
approximation. Similar differences remain for higher-order correlations, derived
by Mo, Ying, and White (1997). They also check both approximations by numerical simulations and observations and find that both work equally well the status
of observational checks does not yet allow us to choose between the two theories.
= ( )
( )
242
2002 by CRC Press LLC
PressSchechter method
231
The PressSchechter theory is meant to provide extensive distributions (different mass functions). It has also been used to describe merger processes in a
hierarchical growth of the observed structure. It ignores many processes that are
important for the growth of the structure (see Monaco 1999), but describes the
mass functions rather well. The main difference with a better founded peak picture is that the PressSchechter theory does not include the notion of a location of
the halo. This has led to an implicit mixed theoretical paradigm of isolated virialized dark matter halos (peaks), which have a PressSchechter type mass function,
and are populated by galaxies.
In recent years, numerical simulations have shown that the standard Press
Schechter function over-estimates the abundance of high mass halos and underestimates those of low mass (Sheth and Tormen 1996), as predicted by Peacock
and Heavens (1990). We shall write the ShethTormen mass function by noting
first that a mass function can be written as
n(M ) d ln M
f ( ) = M 2
= ( )
d ln
for any variable M . The original PressSchechter theory predicts that the
left-hand side of this relation depends only on and for C = M it is:
f ( ) =
2 e
( )
2 =2 :
= C2 =2 (M );
and in this case
f ( ) =
p1 1=2 e
2
(7.62)
=2 :
For the variable (7.62) Sheth and Tormen found from numerical simulations a fit:
1 + ( 0) p ( 0)1=2 e ( )=2 ;
p
where 0R = a . Their best fit has a = 2, p = 0:3, and A is defined by normal= 1. The standard PS function has the same form, if a = 1,
ization, n( ) d p
p = 0, A = 1=(2 2 ). A later work (Sheth, Mo, and Tormen 2001) showed
n( ) = A
that this mass function can be derived by the excursion set approach for elliptical
collapse, where the barrier height is not constant, but depends on the ellipticity
distribution.
The excursion set approach gives also bias functions, using the same methods
as above. The (Eulerian) bias function derived by Sheth and Tormen (1996) is
0
bE (1 ; 1 ) = 1 + 1
1
1+
243
2002 by CRC Press LLC
2p :
1 (1 + (10 )p )
232
()
2. The halos are supposed to have a universal density profile. This assumption is
supported by the latest N -body simulations, and the density profiles proposed
are the NFW profile
(r) =
0
; y r=rc ;
y(1 + y)2
(7.63)
and the M99 profile (see Peacock and Smith 2000 for more details):
(r) = 3=2 0 3=2 ; y r=rc ;
(7.64)
y (1 + y )
where rc is the core radius. The mass of the halo is M = (4=3)v rv3 , where
rv is its outer radius and the virialization density contrast is taken as v = 200
(a popular convention instead of = 178 given by spherical collapse, Chap-
ter 6). Each profile has its advantages the NFW profile has an analytical
Fourier transform (White 2001) and the M99 profile has an exact mass-radius
relation (Peacock and Smith 2000).
( )=
Here M? is the nonbiased halo mass, familiar from the PressSchechter theory
C , see Section 7.7). The parameters of the relation (7.65) are
( M?
( )=
244
2002 by CRC Press LLC
233
10, 0:2
4. The fourth component is the halo mass function. The usual choice here is the
recent modification of the standard PS function by Sheth and Tormen (1996).
We shall follow their convention and redefine the density contrast parameter
2 = 2 M , where C
: and M are defined as in the standard
as C
PressSchechter theory. This mass function n is written as
p
n A
a p ae a=2 ;
( )
= 1 69
( )
()
( )= 1+( )
p
with the standard
R choice a = 2; p = 0:3 and A coming from the normalization condition d = 1.
5. We know that the spatial distribution of massive halos is biased, so one of the
ingredients of the model has to be the bias function. This function is derived
on the basis of the mass function above (Sheth and Tormen 1996) and reads
b( ) = 1 +
a
c
1+
2p
1 + (a )p :
These are all the assumptions needed to describe the clustering of dark matter.
The power spectrum of dark matter Pdm k is represented as a sum of two parts,
one describing the halohalo correlations and the other the short-range correlations within halos. To get the first part we have to take into account that halos are
extended, so we have to convolve the initial density with halo profiles. In Fourier
space this reduces to multiplication with an averaged filter:
()
hh (k) = P (k)
Pdm
lin
Z
2
(7.66)
The only new function in the above formula is the normalized Fourier transform
of the halo density profile
y(k; ) =
( )
e(k)
;
M
y(0) = 1
(the function y k; depends on via the halo mass M ). This formula also introduces a constraint on the bias function (Seljak 2000). Because for large scales
(k ! ) the halo sizes do not matter, the power spectrum has to coincide there
with the linear Plin . This gives
n( )b( ) d = 1;
()
which in combination with the normalization condition for n says that when
halos are biased (b > ) in some mass range, there must be also a mass range
where they are anti-biased (b < ).
245
2002 by CRC Press LLC
234
The second term of the power spectrum comes from correlation of mass within
halos:
Z
M
h
Pdm
(7.67)
n
jy k; j2 d
3
= (21 )
( ) ( ) ( )
()
( )+
( )=
White (2001) extended this model to include redshift space distortions, describing these by the direction-averaged distortion factors. The amplification of the
linear spectrum in the plane-parallel limit is
A1 () = (1 + f2 )2 ;
where is the cosine of the angle between the line-of-sight and the wave direction
and f 0:6 is the dimensionless growth factor; these distortions are explained
in more detail in Chapter 8. The velocity distortions caused by virialized motions
inside a halo suppress the spectrum by
2
A2 (k; ) = e (k) ;
2 =
GM
2rv :
The total effect of the redshift distortions can be written as the product of the two
factors, and the total average redshift distortion is given by an integral
Denoting k
R(k; ) = 12
Z 1
p erf(x)
R(x) = 8 x5
2
e x
2
2 2
2
4x4 [3f + 2f y + 4fy ]:
Using these distortions, the formulae for the redshift space spectrum can be obtained, including the function R k in the integrands in (7.66) and (7.67). This
analytical model is compared with results of numerical simulation in Fig. 7.10.
The power spectra in the figure are given in the power-per-log-interval presenk3 P k = 2 ). Both the real space power spectrum and the
tation ( 2 k
velocity space power spectrum, predicted by the model, describe the simulation
rather well. As the latter is the result of complex nonlinear dynamical evolution,
it suggests that the power spectrum is insensitive to the details of this evolution
and reflects only its initial and final stages.
The halo model would not be very useful, if it could predict only dark matter
distributions. These are certainly extremely important, but in order to check theory
( )
()=
( ) (2 )
246
2002 by CRC Press LLC
235
Figure 7.10 Halo model predictions for real and redshift space power spectra (solid lines,
the line close to the solid squares gives the real space power spectrum). The results of a
N -body simulation with the same initial spectrum are shown by squares. Solid squares
indicate real space power and open squares indicate redshift space power. The dashed
and dotted lines show the halohalo and in-halo contributions to the redshift space power
spectrum. (Reproduced, with permission, from White 2001, Mon. Not. R. Astr. Soc. 321,
13. Blackwell Science Ltd.)
with observations a model must describe galaxy distributions. The halo model can
do that, if we add three more components to the model.
6. To get from halos to galaxies, an essential quantity is the average number of
galaxies per halo, N M . This number should be a function of halo mass; a
: is suggested by N -body
power-law slope N M =M M with
simulations.
( )
( )
02
7. Assuming that galaxies follow mass inside halos, the simplest function to describe the difference between the dark matter and galaxy correlations is the
hN Np i. This also can
mean number of galaxy pairs per halo, NP M
be approximated by a power law; simulations suggest NP M =M M
with the same value of
: as above.
( )= (
02
1)
( )
8. We assume that galaxies follow mass in halos. An important (but natural) assumption of the halo theory is that there is always the central galaxy. If the
number of galaxies in a halo is large, this does not matter; the galaxy and dark
matter correlations are similar, given by the convolution of the density profile.
247
2002 by CRC Press LLC
236
Figure 7.11 Galaxy power spectrum predicted by a halo model (the line marked PgP +
hh
Pg ) compared to the APM galaxy power spectrum (dots with error bars). Different contributions to the galaxy and dark matter power spectra are shown; the notations in the
legends differ only by using the superscript P for the single halo part of the spectra. (Reproduced, with permission, from Seljak 2000, Mon. Not. R. Astr. Soc., 318, 203213. Blackwell Science Ltd.)
If there are only two galaxies in a halo and one of them is in the center, the
correlation function coincides with the density profile itself.
These assumptions modify the expressions for the power spectrum. The halo
halo part is now
n
n( )
2
N (M )
b( )y(k; ) d ;
M
(7.68)
where n is the mean number density of galaxies (in a sample). The factor
comes from the normalization condition
Z
=n
n
N (M )
n( ) d = :
M
Pgh
2 Z
M ( ) NP (M )
1
n( )
jy(k; )jp d;
= (2)3 n
M2
248
(7.69)
( )
237
where the term jy k; jp accounts for the different correlation functions for sparse, if NP M , the correlation funcly populated halos. If NP M > , p
. The total power spectrum is
tion is given by the halo profile and p
Pg k Pghh k Pgh k :
(7.70)
( ) 1 =2
( ) 1
=1
( )= ( )+ ( )
Fig. 7.11 shows an example of the galaxy power spectrum calculated by formulae (7.687.70). It shows both the dark matter and galaxy power spectrum term
by term and compares it with the observationally determined (APM) power spectrum. The convention for the power spectrum is proportional to the 2 measure
used in the previous figure. We see that the halo model reproduces well the positive bias for smaller scales, due to the switch-over in the form of the correlation
function, and is close to the observations for a wide wavenumber interval.
The halo model consists of several components that can be studied and improved separately. For example, Peacock and Smith (2000) calibrated the N M
dependence by using data on the luminosity function of galaxy groups and applied
the halo model to halos found by numerical simulation.
In addition to the description above, they stress that the halo model also dein small halos. If we take into account
scribes isolated galaxies with NP
that galaxies can orbit inside their parent halos, the small-scale biasing has to be
nonlocal. They also show that the halo model can explain the low amplitude of
the pairwise velocity variance of galaxies.
Predictions for galaxy power spectra can also include velocity distortions, as
done above for the dark matter. Seljak (2001) has predicted the redshift space
bias for several types of galaxies, showing that it is also nonmonotonic, as is the
real space bias.
The halo model as delineated above predicts the power spectrum of the galaxy
distribution. Scoccimarro and Sheth (2001) used the halo approach to generate
model galaxy distributions. They replaced the time-consuming N -body modeling
by approximate second-order Lagrangian perturbation theory predictions, used
the halo merger tree algorithm of Sheth and Lemson (1999) to select virialized
dark matter halos, and generated galaxies inside these halos according to the halo
model rules. Their method is similar to the peakpatch method, but faster, and
reproduces well the features of N -body halos. This method could be especially
useful for maximum-likelihood calculations, when models have to be generated
for a large parameter space.
( )
=0
238
field g
(7.71)
where b1 is a constant bias factor. Dekel and Lahav (1999) have generalized this
simple model. They note first that the common result b > together with (7.71)
cannot be applied for voids where it could predict negative galaxy densities. Second, they stress that there are certainly many other variables influencing galaxy
formation other than the local overdensity alone. This leads to a stochastic biasing
model where both and g are random fields and the biasing relation is determined
by the biasing conditional distribution P g j for a given . The biasing relation
(7.71) will be replaced then by the conditional mean
( )
b( ) = hgj i =
P (gj )g dg:
(7.72)
() as
b = hb()2i=2;
= hb2 ()2i=2:
=g
hgji:
(7.73)
(g) and q() are any functions of g and , we get the average biasing
where p
scatter
Thus we have three natural biasing parameters instead of one the mean biasing b, the (second order) nonlinearity eb=b, and the scatter b =b. These parameters
are illustrated in Fig. 7.12, which is the result of a semi-analytic model of galaxy
formation. In this model both the matter density and galaxy density fields are generated, the latter based on rather complicated empirical prescriptions for galaxy
formation. We see that the model biasing relation is clearly nonlinear, has a pretty
good scatter, and is different for different epochs.
250
2002 by CRC Press LLC
239
Figure 7.12 Biasing of the galactic halo field versus the overdensity field in a N -body
simulation. The left panel corresponds to the present moment (z = 0), the right panel to
an earlier epoch (z = 1). The mean biasing relation is given by a solid curve and its scatter
by error bars. (Reproduced, with permission, from Dekel and Lahav 1999, Astrophys. J.,
520, 2434. AAS.)
Because different methods for estimating bias use different regressions, they
measure different combinations of bias parameters. The most often used bias
value is the ratio of variances
2
b2var = g2
= eb2 + b2 ;
which can be obtained by using the averaging rule (7.73). We also can write that
relation as
!
e
b2 b2
bvar b 2
2 :
= b + b
This shows that bvar is biased compared to the mean bias b. The galaxydensity
covariance can be calculated to be
hgi = b2
b = eb = b1 ;
bvar = b1
2
1 + b2b
1
and for nonlinear and deterministic biasing (as that derived for massive halos)
e
b
6= b;
b = 0;
251
2002 by CRC Press LLC
bvar = eb:
240
Dekel and Lahav (1999) analyze different bias estimates and show that they
determine different combinations of bias parameters. They also extend their formalism to correlation functions and power spectra and to redshift distortions. The
natural assumptions of nonlinearity and stochasticity of bias lead to more complicated relations between the descriptors of the galaxy distributions and descriptors
of the total matter.
252
2002 by CRC Press LLC
CHAPTER 8
E e(k)e? (k0 )
(k)
Here e
is the Fourier amplitude of the density contrast field at a wavenumber
, a star denotes complex conjugation, and E fg denotes expectation values over
is the Fourier transrealizations of the random field. The power spectrum P
of the field, D
is the three-dimensional
form of the correlation function
Dirac delta function, and the factor 3 comes from our Fourier transform convention:
Z
e
f
f eikx d3 x;
(r)
(2 )
(k) = (x)
Z
1
e
f (x) =
(2)3 f (k)e
(x)
(k)
ikx d3 k
(2 )
(we can consider the volume element in the Fourier space d3 k= 3 ). Although
we can see different Fourier transform conventions in cosmological literature, this
is the most common one.
Estimation of power spectra from observations is a rather difficult task. Up to
now the problem has been the scarcity of data; in the near future we will have
the opposite problem of managing huge data sets. The development of statistical
techniques here has been motivated largely by the analysis of CMB power spectra,
where better data were obtained first, and has been parallel to that recently. We
shall see that one of the possibilities to compress the information for large data
sets is to calculate the correlation function as an intermediate step on the way to
the power spectrum.
253
2002 by CRC Press LLC
242
n(x) =
where D
(x)
D (x xi );
(8.1)
hn(x)i (angle brackets denote expectations over the point process) itself is a re-
(x)
(x) =
(x)
(x) (x)
;
(x)
(8.2)
(x) =
(x) (x)
(x x ) 1
(x )
(k) =
(x)(x)eikx d3x
254
(k):
(8.4)
243
(k) =
(x)eikx d3x
V
depends only on the sample geometry and on the weight function.
Using (8.4, 8.3) we get the estimator for a Fourier amplitude
F (ki ) =
(xj ) eikix
n (xj )
(ki )
(8.5)
(8.6)
E fhF (ki )
F?
(ki )ig =
Z Z
x0 ) d3 x d3 x0
(8.7)
Here E fg denotes expectations over realizations of the density contrast field, and
angle brackets denote expectations over the point process.
The last term in (8.7) is obtained by substituting the zero-mean condition for
the density contrast
n
E e(ki )
= 0;
Z
(x) (x)
eiki x d3 x
= (k )
e i :
n
V
We have shown in (3.9) that the second-order intensity of an inhomogeneous
Poisson process is
(8.8)
The first term comes from the independence of the process in separate points, and
the second term accounts for self-pairs there is always a point of the process at
zero distance. When calculating correlation functions, we discard self-pairs. Here
they are usually included, as we sum over all points first to get the estimators of
Fourier amplitudes and take a square later.
The definition of the density contrast (8.2) allows us to write
255
2002 by CRC Press LLC
244
(x)g = n (x). Let us expand now the expectation value in the integral
where E f
(8.7):
E fhn(x)n(x0 )ig = n (x)n(x0 ) 1 + E f (x) (x0 )g + D (x
x0)n(x): (8.9)
Integrating the first term in the square brackets, we get a product of two integrals
that gives j e i j2 , canceling the last term in (8.7). Integrating the third term in
(8.9), we get the Poisson (shot) noise term
(k )
S=
x) d3x:
2(
(x)
V n
(8.10)
V V
Using the convolution theorem, we get
Z
(x)(x)eikix d3 x =
e?
x0 ) d3 x d3 x0 :
3 0
d3 k0 d3 k00
(2)6
3 0
G(ki
It is easy to see that this function is different from zero roughly in the region
k < 1=L , where the index denotes spatial components of a vector. For larger
hjF (ki
)j2 i =
G(ki
3 0
k0)P (k0) (2d k)3
x) d3x;
2(
(x)
V n
(8.12)
or, symbolically, we can obtain the estimate of the power spectra Pb by inverting
the integral equation
G
Pb PR S;
(8.13)
where
denotes convolution, PR is the raw estimate of power (8.6), and S is
the (constant) shot noise term. In order to get an unbiased estimate, the window
has to satisfy the condition
function G
(k)
G(k)
d3 k
(2)3
256
2002 by CRC Press LLC
= 1;
245
x) d3x = 1
2(
(8.14)
by Parsevals theorem.
For small scales (k =L) and in the case where the power spectrum can be
constant throughout the width
expected to be fairly smooth, we can consider P
of the window G . As the window is normalized, the estimate of the spectrum
in that case is
Pb PR S:
In general, we have to deconvolve the noise-corrected raw power to get the estimate of the power spectrum. This introduces correlations in the estimated amplitudes, so they are not statistically orthogonal any more. A sample of a characteristic spatial size L creates a window function of width of k =L, correlating
estimates of spectra at that wavenumber interval.
Fig. 8.1 shows an example of two window functions and a typical density power
spectrum. Both window functions are rather sharp and describe deep samples. The
window on the left is a typical window for a volume-limited sample. Such samples
have sharp boundaries in real space, which cause wide sidelobes in Fourier space.
The window on the right is typical for a pencil-beam survey. A pencil-beam survey
is a deep survey in a very small region of the sky, and it cuts out a narrow beam in
space. Such surveys have an anisotropic geometry that gives anisotropic window
functions in Fourier space, where the largest extent of the window is determined
by the smallest linear size of the sample. A deep survey in a thin slice of the sky
has a similar Fourier space window function.
As we see, a wide window function will mix up the signals from different
scales, severely limiting the resolution of the spectra we obtain. This problem
is especially serious for small wavenumbers that are comparable with the width
of the window, and also for regions where the amplitude of the spectrum is small.
There the leakage of high-amplitude power can seriously distort the estimated
signal.
(k)
(k)
246
Figure 8.1 A standard (CDM) density power spectrum (center) and two window functions. The window function on the left is typical for volume-limited
samples, and that on the right is typical for low-dimensional surveys (e.g., for a pencilbeam survey). Both the power spectrum and the window functions are three dimensional;
one dimension is suppressed in the figure. (Reproduced, with permission, from Tegmark
1995, Astrophys. J., 455, 429438. AAS.)
F (k) are Gaussian variables, Feldman, Kaiser, and Peacock (1994) found
E fPb(k)Pb(k0 )g = jE fF (k)F ? (k0 )gj2 :
(8.15)
The assumption of Gaussianity is certainly justified for scales that follow linear
dynamics.
The covariance of the raw estimates in (8.15) can be found in the same way that
brought us to their variance in (8.12). This gives
Z
F?
(kj )ig =
G(ki
imately constant within the width of the window function. Using the convolution
258
2002 by CRC Press LLC
247
2 ( )ei(kj ki )x P
= (k + k ) 2
k k 1
+ n(1x) d3 x;
(8.17)
where P P
i
j = . Because of the oscillating factor in this integral, raw
estimates of spectrum are uncorrelated for j i
j j =L and well correlated
for j i
j j =L. This leads to the notion of coherence cell in the wavenumber
space; estimates of raw power spectra are independent only for different coherence cells.
The formula for the covariance of the power spectrum (8.15) gives for the variance
P2 i jPR i j2 :
Thus the rms error of a single power spectrum estimate (or the estimate for a single
coherence cell) is equal to the raw spectrum; this estimate is extremely noisy. The
only way to reduce the noise is to average over several coherence cells. As the
cosmological spectra are usually assumed to be isotropic, the standard method to
estimate the spectrum involves an additional step of averaging the estimates Pb
over a spherical shell k 2 ki ; ki+1 of thickness ki+1 ki > k
=L in the
wavenumber space. The variance of this estimate is
k k
(k ) =
(k )
=1
(k)
Z Z
3
3
2
2
E fhF ( i )F ? ( j )ig2 d ki d kj
P (k ) = 2
VS VS VS
(2)3 (2)3
(the factor 2 is due to the fact that the density field is real and only half of the
coherence cells are independent). If the width of the shell is large compared to
the coherence length, one of the integrals gives the volume of the shell. Using the
result (8.17) and Parsevals theorem, we get
P2 (k) =
2P 2 (k) Z
x) 1 + n(x)1P (k)
2(
2
d3 r:
(8.18)
VS
The optimal weight function will give the minimum variance for the power spectrum estimates. Minimizing (8.18) together with the normalization condition (8.14)
gives
VS
(8.19)
This weight function was derived first by Feldman, Kaiser, and Peacock (1994).
Another derivation of this result is given by Tegmark et al. (1998); they also show
that this weight function is optimal only for k =L.
This weight function is, in principle, different for every wavenumber and demands a priori knowledge of the amplitudes of the power spectrum. However, if
the weight function is normalized according to (8.14), the estimates will remain
unbiased. An error in the value of P will only increase the variance of the estimates. Hence in applications a few constant values of P are usually used.
259
2002 by CRC Press LLC
248
where Vs
as:
P2 (k)
Pb2 (k)
V (k)2V (k) ;
s
(k) is the volume of the shell in the wavenumber space and Ve is defined
Z
n (x)P (k) 2 3
Ve (k) =
(x)P (k) d x:
V 1+n
As the integrand is about one in regions where the signal dominates the noise,
P =n, and zero where the noise dominates, this volume is called the effective
volume. For a volume-limited survey (n is constant) we get
1
Ve (k) = V
1 + nP1(k)
=1
()
( )
Pb2
() 2
(k) N
where N is the number of coherence volumes in the wavenumber space shell. The
number of independent volumes is twice as small (the density field is real). We recall that the above results were obtained assuming that the probability distribution
of spectral amplitudes is Gaussian.
It is instructive to compare the behavior of the estimates of the correlation function and the power spectrum, if we vary either the number of galaxies in the sample (the mean density n) or the size of the sample L. If we observe more galaxies
within the same volume, the variance of the correlation function decreases. Alternatively, we can increase the resolution of the correlation function, keeping
the same variance level. We also can estimate the correlation function at smaller
separations.
The main effect of a larger sample density on the power spectrum estimate
is to reduce the shot noise level; the variance of the estimate practically does
not change (it decreases only slightly due to the reduced contribution by shot
noise). Thus, measuring positions of more galaxies in the same volume, we do not
improve the estimate of the power spectrum. A sample of higher density allows us
to extend the wavenumber space (the maximum wavenumber, for which we can
expect to have reliable power spectrum estimates, is given by k =d, where d
is the average separation between galaxies).
If we increase the sample depth L, the variance of the correlation function
estimate decreases; there are more pairs for a particular bin. As above, we can
increase the resolution of the correlation function, but we cannot decrease the
260
2002 by CRC Press LLC
249
lower limit of pair separations. Naturally, we can estimate the correlation function
for larger pair separations.
The variance of the estimate of the power spectrum for a coherence cell of
width k =L does not depend on the depth L. Increasing the depth of the
survey will shrink the coherence cells and increase the resolution of the power
spectrum. The variance of the power spectrum will decrease only if we decide
to keep the same final resolution; the final estimates will then be averaged over
more coherence cells. A larger depth of a sample will also allow us to extend the
spectral range to smaller wavenumbers.
The FKP weight function (8.19) has been the standard almost from the time of
its derivation by Feldman, Kaiser, and Peacock (1994). Other methods that have
been used include pure volume weighting (the weighting function is constant inn that weights all objects
side the survey volume) and a function
equally. The FKP weight function uses volume weighting for regions where the
mean density is large enough and the density is well sampled. For smaller densities (near the boundaries of a sample) it switches over to give equal weight to
individual objects.
These weight functions are appropriate for estimating power spectra at large
wavenumbers, k =L. For smaller wavenumbers (large scales), where the window width is comparable to the wavenumber, the careful choice of proper weighting will be crucial. Tegmark (1995) has shown that the optimal selection function
is the eigenfunction of the equation
(x) = (x)
(x)
r2 n ( x) (x) = E (x)
for the minimum eigenvalue E . The constant
is a free parameter that gives the
ratio of the rms error of the power spectrum estimates to the resolution in the
wavenumbers (the ratio of vertical error bars to horizontal error bars). The
describes also the geometry of a sample, as n
selection function n
outside sample boundaries.
(x)
(x) = 0
261
2002 by CRC Press LLC
250
(the density profile is easier to determine than the mean density). This would
change the expectation of a Fourier amplitude estimate from zero to
E fF (ki )g = (a
1) e(ki )
and the spectrum would be shifted by the square of this amount. Because we
do not know the true mean density and a, the solution is to modify the weight
function to make its Fourier transform vanish at i .
The integral constraint
n(x)
a
n 0 (x)
V
Z
1 (x) d3 x = 0
Z
1
1 X (xj ) :
n(x)
A= e
(
x
)
d3 x = e
(0) V n0 (x)
(0) j n0 (xj )
Substituting that into (8.4) we can write, after a few intermediary steps,
(k ) =
e i
where
(x) =
a
A
n(x)
(x) d3 x;
n (x) i
"
eiki x
(ki) (x)
e(0)
(k) = e(ki + k)
(0) = 0
(8.20)
a=A
(8.21)
1 in (8.20)). Its
(ki) e(k);
(0)
251
The new weight functions change, however, the shot noise term and the window
amplitude. The window integrates now to
Wi =
Z
iki x
e
(ki) 2 d3 r:
e(0)
Ni =
Z
iki x
e
G
Pb
PR (ki )
=
W
i
Ni
i
Different approximations have been used to calculate the integral constraint
correction in the past. Tegmark et al. (1998) show that these usually give larger
corrections than needed, leading to lower power spectrum values at small k .
As an example of the application of the direct method we show the power spectrum obtained by Miller and Batuski (2001) for a carefully selected deep sample
of rich Abell/ACO clusters of galaxies. The total number of clusters in the sample
is 545, and the sample geometry is typical a double cone due to obscuration
by dust near the plane of the Galaxy. In this sample the clusters with galactic
; were excluded. Also, observational projects needed to
latitude b 2
carry out full-sky surveys have to use several different instruments, and so frequently the survey depth is different in different directions in the sky. For this
survey the northern galactic cone has a depth of 420 h 1 Mpc, and the southern
cone has a depth of 300 h 1 Mpc. The spectrum is shown in Fig. 8.2 together
with several other spectra obtained for smaller cluster samples. We see that the
spectrum is rather close to a power law with a probable plateau (or maximum)
around k : h Mpc 1 . The reality of that maximum is not clear, however.
The hatched region in the figure shows where the width of the window function is
comparable with the wavenumber; thus the estimates in this region are inaccurate.
But these estimates indicate that the spectrum is yet rising toward the wavenumber
k
: h Mpc 1 . The spectra for smaller samples shown in the figure tend to
decrease at the maximum scales of a sample; this could be the result of insufficient
correction for the integral constraint in these studies.
The deepest galaxy sample obtained thus far is the ongoing 2dF redshift survey, described in Chapter 1. The power spectrum for this survey was estimated
by a careful application of the direct method by Percival et al. (2001). The survey
contained then about 160,000 galaxies and had a depth of 750 h 1 Mpc, deeper
than the cluster survey described above. Fig. 8.3 shows the power spectrum of that
survey, normalized to a model (CDM) power spectrum. This spectrum is a convolution of the true power spectrum with the survey window function; convolution
[ 30 30 ]
0 02
= 0 01
263
2002 by CRC Press LLC
252
Figure 8.2 Power spectrum of Abell/ACO clusters (solid circles) compared with those
found for smaller samples. Open circles show the spectrum for very rich clusters in the
same catalog, triangles show the spectrum for a smaller Abell/ACO sample, and stars
show the spectrum for the APM cluster sample. Estimates of the power spectrum in the
hatched region are inaccurate. (Reproduced, with permission, from Miller and Batuski
2001, Astrophys. J., 551, 635642. AAS.)
0 02
may distort the results at wavenumbers k < : h Mpc 1 . The distortions of the
power spectrum due to peculiar velocities and nonlinear processes were estimated
using large-scale N -body simulations; these distortions compensate one another
for k < : h Mpc 1 . The region where the estimates of the power spectrum are
reliable is constrained by vertical dotted lines. The error bars were determined
from realizations of a Gaussian Cox process with a similar power spectrum in a
similar survey volume.
0 15
The most interesting features of the 2dF survey power spectrum are its details
(wiggles), which have been interpreted as traces of acoustic oscillations in the
post-recombination power spectrum. Similar features were also found recently
by Miller, Nichol, and Batuski (2001) on the basis of the Abell cluster sample,
and by Silberman et al. (2001) on the basis of peculiar velocity data. Percival
et al. (2001) found that the parameters of the best-fit cosmological model are
: ; b= M
: , which gives a rather high baryon fraction b .
Mh
This concurs with recent results obtained from the CMB data.
= 0 21
= 0 15
264
2002 by CRC Press LLC
253
Figure 8.3 Power spectrum of the 2dF redshift survey, divided by a linear-theory CDM
power spectrum. The spectrum is not deconvolved. Error bars are determined from Gaussian realizations; the dotted lines show the wavenumber region that is free of the influence
of the window function and of the radial velocity distortions and nonlinear effects. (Reproduced, with permission, from Percival et al., Mon. Not. R. Astr. Soc., in press. Blackwell
Science Ltd.)
p=
=1
(d; p)
d=
=1
265
2002 by CRC Press LLC
254
L = ln L(d; p):
(d; p)
P (pjd; I ) = P (pjI )
P (djp; I )
;
P (djI )
(8.22)
where I denotes the prior information. The first factor on the right-hand side of
(8.22) is the Bayesian prior that includes our prior knowledge about the values
of the parameters . The nominator is the likelihood function (note that it can
also depend on the prior information, in principle) and the denominator is the
normalization constant (global likelihood). Confidence values in parameter space
are defined by integrals of the likelihood function over subvolumes of that space.
The normalization factor (global likelihood) is frequently difficult to estimate; in
this case the ratio of integrals of the likelihood over subvolumes is used to find the
odds that one subvolume of parameter space is more likely than the other in the
light of the given data set. If we are interested only in a subset of model parameter,
confidence values can be found by integrating over the remaining parameters (this
is called marginalization).
The problems usually associated with the Bayesian approach, the need to define the prior distribution jI and difficulties in calculating the global likelihood, really stress the need to define exactly the statistical model we compare
with observations. The real meaning of more general frequentist statistics is
usually revealed only after extensive studies. For an introduction to the Bayesian
approach in astronomy we recommend the review by Loredo (1990), which can
be found together with more material at the Web site (BIPS).
In a Gaussian case the log-likelihood function
(p )
(8.23)
and
are the model-predicted mean of data and the data covariance
where
T ? means the Hermitian conjugate of the matrix
matrix. The notation y
(the complex conjugate of its transpose).
It would be difficult to build a model that would predict the coordinates of the
observed objects. The power spectrum describes a continuous field ; thus a
A = (A )
(x)
266
2002 by CRC Press LLC
255
di =
(x)
Z
n(x)
n (x)
(x) d3 x;
(8.24)
(x)
(x)
(x) = (x)exp( k x)
( )
Fij = E
@2L
:
@pi @pj
A thorough analysis and examples of the Fisher matrix are given by Tegmark,
1 can be thought
Taylor, and Heavens (1997). The inverse of the Fisher matrix
of as the covariance matrix of parameter errors. It can be strictly proved that,
267
2002 by CRC Press LLC
256
=
=(
= exp( )
L
(T 1 )ij = @p@ @p
2
The Fisher matrix is the expectation value of these coefficients, and its inverse
1 can be used as the covariance matrix of the parameter estimates.
In order to find the confidence regions for the parameters, we should integrate
over the parameter space; to find error estimates for single parameters, we should
marginalize over the parameter distribution, which involves more integrals. These
multidimensional integrals are usually difficult to calculate. Using the Fisher matrix, we do not have to worry about integrals. Nevertheless, we have to keep in
mind that this is essentially an approximate method (that works well most of the
time).
As the Fisher matrix is an expectation value over realizations of a random field,
it depends only on the statistical model. For a real multivariate Gaussian distribution, for example, the Fisher matrix can be calculated as
Fij =
(8.25)
where the commas denote derivatives with respect to model parameters, (C;i )kl =
@Ckl =@pi . So the Fisher matrix can be calculated beforehand to define the observational strategy and to evaluate the model.
There are two popular choices of parameters to describe the density power
spectrum in cosmology. One is a set of about 7 to 8 parameters describing the initial simple power spectrum (its amplitude and exponent), the number of neutrino
flavors, the main cosmological parameters ( M ; ; baryon ; H0 ), etc. Another
choice is to describe the power spectrum by band-powers, its values at certain
wavelengths, approximating the power spectrum by a step function. In the latter
case the elements of the Fisher matrix give directly the covariance of the estimates
of the power spectrum.
268
2002 by CRC Press LLC
257
d=
d
d
di i ;
i
where i form an orthonormal basis in the space of data vectors. If we rotate the
basis vectors to a new system 0i , the new coordinates of the data vector in this
0i ? (the new basis is also orthonormal,
0?
basis will be d0i
0i ? 0j ijK , where ijK iis the Kronecker
delta).
As the noise is usually statistically independent of the signal, the total data covariance matrix can be written as a sum of the signal and noise covariance
matrices,
:
(8.26)
We shall suppose first that we can neglect the noise. The (model) data covariance
matrix
Sij E di d?j
will be transformed to
0 y E di d? 0 :
Sij0
i
j j
Demanding that the new covariance matrix 0 is diagonal,
(e ) e =
= d (e ) = (e ) d
C=S+N
=
= (e )
S0
ij
we get
= iij ;
(8.27)
Se0i = ie0i:
This means that the new basis vectors are the eigenvectors of the covariance ma0ii . This coordinate transformation is called the distrix with eigenvalues i
crete KarhunenLo`ewe (K-L) transformation. Note that the covariance matrix
is the model covariance matrix that depends only on the proposed spectrum (it is
also called the fiducial spectrum) and pixelization.
The condition (8.27) tells us that the new data values are statistically orthogonal. If the data has zero mean, then they are also uncorrelated, and if the data
=S
269
2002 by CRC Press LLC
258
distribution is Gaussian, the new data values are statistically independent. All of
these conditions are satisfied most of the time in cosmological problems.
Vogeley and Szalay (1996) show that the K-L transform is unique and that this
basis is optimal to represent the data. If we arrange the modes by decreasing
and select the first m < n components of the new data vector, then the K-L basis
gives the minimum rms truncation error.
8.2.6 Signal-to-noise eigenmodes
N= I
2 , where is the
If the noise is constant per original data point (pixel),
unit matrix, then the K-L transform, a coordinate rotation in the data space, leaves
the noise matrix diagonal and the signal and noise remain uncorrelated.
If the noise per pixel is not constant, the noise matrix is diagonal, but not proportional to the unit matrix, and diagonalization of the signal covariance matrix
will generate a nondiagonal noise covariance matrix. The remedy for that is to
rescale the coordinates first to transform the noise covariance matrix to the unit
matrix,
N00 = WyNW = I:
We have written the formula above to be usable for more general transformations
diag N1 ; : : : ; Nn ,
than simple rescaling. In the case of
p rescaling, ifpwe write
diag = N1 ; : : : ; = Nn . Vogeley and Szalay (1996)
the scaling matrix
prewhitening, as the scaled data vector
call the coordinate transformation by
has white noise. The covariance matrix for the prewhitened data is
W=
(1
N=
Using
WWy = N
1 we get
(8.28)
N S
1 , they
As the K-L eigenvectors of prewhitened data are the eigenvectors of
also are frequently called the signal-to-noise eigenmodes. We also can write
(8.28) as
0 i 0 ;
(8.29)
i
i
redefining the eigenvalues . Thus, the K-L eigenvectors can also be found as the
solutions of the generalized eigenvalue problem (8.29). If the noise for different
can be more general,
K-L modes is correlated, the prewhitening transform
including both rescaling and rotation of the coordinates. The derivation above
remains in force for that case, too.
These modes are optimized with respect to the signal-to-noise ratio, and those
are the modes commonly used to estimate the power spectra. We also can find
Se = Ne
270
2002 by CRC Press LLC
259
other K-L modes, optimized with respect to different parameters of the covariance
matrix. As shown by Tegmark, Taylor, and Heavens (1997), the eigenvectors for
this case can be obtained from
@C 0
e = iCe0i:
@pi i
(8.30)
The formula (8.29) above is a special case of (8.30) for the parameterization
Heavens (1997) also show how to combine the representations of data, optimized
for different parameters.
The total number of the K-L eigenmodes is the same as the length of the data
vector (n) because they represent the data vector in the same space. As the rms
amplitude of the noise is constant (unity for prewhitened data), the eigenvalues
can be thought of as the signal-to-noise ratios. Usually many of the K-L modes
have eigenvalues much less than one and we can truncate the total data vector by
retaining only the modes with high eigenvalues. As the size of the covariance matrices is n2 , this compression frequently makes the likelihood problem solvable.
As an application of the K-L transform we show you the eigenfunctions found
for the famous CfA2 slice of galaxies by Vogeley and Szalay (1996), who give
a detailed pedagogical treatment of the K-L process in this paper.
The CfA2 slice was the first sufficiently deep sample of galaxies where the
peculiarities of the galaxy distribution were clearly seen (de Lapparent, Geller,
and Huchra 1986). The maximum extent of the sample on the sky is 117 , in
the perpendicular direction 6 (from h to h in RA and from : to :
in declination), and its depth is 150 h 1 Mpc, so it is really a wedge-like slice in
space. Because it is so thin, we shall take it as twodimensional for the present
example. Fig. 8.4 shows the positions of galaxies in this sample.
Vogeley and Szalay binned this galaxy distribution, prewhitened the data vector, and computed the K-L eigenmodes, using the known selection function and
the power spectrum. The K-L transform also can be thought of as the selection of
a new pixelization with the weight functions
17
26 5
32 5
(8.31)
260
Figure 8.4 Two-dimensional projection of galaxy positions in the CfA2 north sample, the
CfA2 slice (see Fig. 1.11). (Reproduced, with permission, from de Lapparent, Geller, and
Huchra 1986, Astrophys. J. Lett., 302, L1L5. AAS.)
10
11
12
Figure 8.5 The most significant 12 KarhunenLo`ewe eigenmodes for the CfA2 north slice.
The gray density is proportional to the mode amplitude. (Reproduced, with permission,
from Vogeley and Szalay 1996, Astrophys. J., 465, 3453. AAS.)
272
2002 by CRC Press LLC
261
Binned
Truncated n=500
Error
Figure 8.6 Demonstration of the K-L transform. The top panel shows the binned density
distribution for the CfA2 slice, the middle panel shows its reconstruction by the 500 most
significant K-L modes, and the bottom panel shows the truncation error. The gray density
is proportional to the number density in bins. (Reproduced, with permission, from Vogeley
and Szalay 1996, Astrophys. J., 465, 3453. AAS.)
273
2002 by CRC Press LLC
262
We can choose the original weight functions to satisfy the integral constraint:
E fdi g = 0:
(8.32)
This is easy to do for pixelized data, as we shall show below. By the derivation
similar to that for the expectation of power amplitudes above we can get the expression for the signal covariance matrix :
S
d3 k
Sij = ei (k) ej? (k)P (k)
(2)3 :
Z
(8.33)
i = Sii =
Gi (k)P (k)
d3 k
(2)3 ;
(8.34)
Gi (k) = ei (k) :
The noise covariance matrix is a unit matrix now, so
Cii = i = wi Pi + 1;
where Pi is the mean power at the wavenumber interval determined by the win(the band-power). Prewhitening means that this window is
dow function Gi
W W y ii . For simple rescaling and a connot normalized to unity, but to wi
n
stant initial weight for all pixels in (8.24) the window normalization is wi
and the K-L band-power estimator is
(k)
=(
Pi = i
n
=
1:
The K-L band-powers are not meant to directly estimate the power spectrum,
because they are not optimized for resolution in the wavenumber space. The spectral windows for the most significant modes tend to be rather wide, and the overlap
between the modes is substantial (see the examples in Vogeley and Szalay 1996).
The calculation of the model covariance matrix is easy for the parameters that
we used to define the K-L transform; it is diagonal in the basis formed by the
K-L eigenmodes. When we minimize the log-likelihood function, the covariance
matrix does not remain exactly diagonal any more, but it will be sparse and easier
to handle than a dense covariance matrix.
The K-L modes depend both on the properties of the sample (geometry and
the selection function) and on the fiducial power spectrum, chosen a priori. As
these modes constitute an invertible transformation of the original data pixels, the
choice of a fiducial spectrum does not bias the final estimates of the power spectrum. As tested in practice, a seriously incorrect choice of the fiducial spectrum
can only slightly enlarge the errors of the estimates.
274
2002 by CRC Press LLC
263
=0
= ()
C=
C = C
where the derivative ;i @ =@pi of the covariance matrix with respect to the
band-power can be computed as
(C;i )ab =
ki jkj<ki+1
ea
C =N
=1
d
q
+1
1
qi = dy C 1 C;i C 1 d:
2
The Fisher information matrix for a Gaussian d and m = 0 is (see 8.25)
Fij = 21 C 1C;i C 1C;j :
Tegmark (1997) showed that the band-powers can be found directly, without
using a costly maximum-likelihood procedure. The mean and covariance of the
compressed data set are determined by the Fisher matrix:
E fqg = Fp;
E fqgE fqgT
(8.35)
= F:
Let us separate now the shot noise term. Formula (8.35) gives
m
X
E fqi g Fi0
Fij pj
j =1
275
2002 by CRC Press LLC
(8.36)
264
=1
(the parameter p0
). So we can restrict the data and parameter vectors and the
; : : : ; m and write
Fisher matrix to the index range i; j
=1
E fqg = Fp + f ;
where the shot noise contribution f is the first column of the original Fisher matrix, fi Fi0 . Generalizing the simple unbiased estimator (see 8.35)
pb = F 1q
(8.37)
pb = M(q f ):
W = MF
pp
b bT
(8.38)
E fpb g = Wp;
E fpb gE fpb gT = MFMT ;
(8.39)
(8.40)
M=F
=1
pbi =
1 Fp = 1 E fqgi;
i a
ai
i
the window-scaled quadratic amplitudes. This estimate is guaranteed to be positive and has minimum variance, although the errors of the amplitudes are correlated.
If we wish to use the band-powers for further maximum likelihood analysis,
correlated errors make our work difficult. Fortunately, there exists a possibility to
1=2
get band-power estimates with uncorrelated errors. This is the choice
(Hamilton and Tegmark 2000) that makes the covariance matrix of the parameter
in this case is
estimates (8.40) diagonal. An exact expression for
M=F
Mij =
"
m
X
F 1=2
# 1
F 1=2
(8.41)
;
ik
ij
k=1
where the normalization in (8.41) is caused by the condition that the rows of the
276
2002 by CRC Press LLC
265
Figure 8.7 The window functions for the 130 h 1 Mpc depth volume-limited CfA/SSRS
sample using the decorrelation method. The windows have been rescaled to the inverse
variance of the band-power estimate. (Reproduced, with permission, from Padmanabhan,
Tegmark, and Hamilton 2000, Astrophys. J., 550, 5264. AAS.)
W = MF
window matrix
have to sum to unity. The window functions obtained
by Padmanabhan, Tegmark, and Hamilton (2000) for this case are illustrated in
Fig. 8.7. They are highly peaked at the ki where the band-powers were chosen and
overlap little. The real window functions have to be normalized to a unit integral;
in this figure they have been normalized to = 2 pi to show the relative variance
of the band-power estimates.
Compression to band-powers is lossless; they contain all the information about
the power spectrum (Tegmark 1977). Thus the parameters of the model power
spectrum can be found by maximum-likelihood (or Bayesian) methods, using
band-powers as the data vector. As the number of band-powers is much smaller
than the size of the original data vector, the gain in speed is appreciable.
1 ( )
mi =
(x)d3 x:
m can be written as
Because we do not know its right normalization, we can choose coordinates that
do not depend on it. In a vector space these coordinates span the subspace per277
2002 by CRC Press LLC
266
= I m(mym) 1my:
The projected data vector is d0 = d, E fd0 g will be identical to zero, and the
new covariance matrix will be C0 = Cy . The new data vector is isolated
from possible uncertainties in the mean density. This, however, comes at a price;
the new covariance matrix is singular. If we use the maximum likelihood method,
1 by the
we need to invert that. Tegmark et al. (1988) recommend replacing
1
y
pseudo-inverse
. The result does not depend on
.
If we estimate the power spectrum using quadratic compression, there exists
a more convenient choice of the projection (Tegmark et al. 1998). The optimal
(minimum-variance) projection operator is
C + mm
1:
d = d
278
2002 by CRC Press LLC
Redshift distortions
267
Figure 8.8 The redshift space view of a evolving spherical overdensity in real space. The
panels show the redshift space density contours in an equatorial plane of the sphere. The
left panel shows a sphere that is far from the observer; the right panel shows a nearby
sphere. (Reproduced, with permission, from Hamilton 1998, in The Evolving Universe,
Kluwer Academic Publishers.)
s = r + br v=H0;
=r v
where r is the distance of the galaxy in real space, v b is the peculiar velocity
of the galaxy (the projection of the galaxys velocity on the direction from
the observer to the galaxy b), and H0 is the present Hubbles constant. We shall
consider nearby samples where the geometry of the space is Euclidean and the
change of Hubbles constant with redshift can be ignored. We have also assumed
that the observer is at rest. We can define that with respect to the CMB rest frame,
subtracting our (well-known) velocity with respect to that frame from the radial
velocity of the galaxy.
The distance distortion cannot change the number of galaxies, so
ns (s)d3 s = n(x)d3 x;
where ns and n are the number densities of galaxies in redshift space and in real
space, respectively. Introducing the density contrast, the above relation can be
written as
n (s)[1 + s (s)] s2 ds d
= n (r)[1 + (r)] r2 dr d
;
279
268
where s are are the density contrasts in redshift space and in real space, respectively, and d is the appropriate region on the sky. The average densities ns
and n are the selection functions of the sample and are the same to the first order. More accurate treatment of the selection functions can be found in Hamilton
(1998). Let us discard the nearby regions of the sample. The peculiar velocities
then satisfy the condition v= H0 r . We also assume that (the linear
approximation that describes large-scale dynamics rather well). Then we get
(s)
(r)
1 + (s) =
s
n (r)
n (r + brv=H0 )
This gives
1 + Hv r
0
2
1 + H1
@v
0 @r
1
[1 + (r)] :
(s) = (r)
2 + n(rr) dndr(r) Hv r :
0
Introducing the dimensionless derivative of the selection function (r) by
d ln(r2 n (r))
(r) = 2 +
;
d ln(r)
@v
H0 @r
(8.42)
(r) v
+
:
(r) = (r) @v
@r
r H0
We have used here the fact that in the first order s (s) = s (r).
The velocity v is connected with the density contrast by the equation of conwe can write
tinuity. In the linear approximation this connection is given by the formula (see
Chapter 6):
H a f a M r ;
(8.43)
where H a is the Hubble function and f a is the dimensionless growth rate of
the growing mode:
()() = v
()
d ln D1 (a)
f (a) =
:
d ln(a)
()
We have written M in (8.43) to stress that it describes the total density. The
galaxy density contrast G is, in the first approximation, proportional to that:
G = bM ;
where b is the biasing factor. Recalling the fact that the growing mode of the
velocity in the linear approximation is irrotational, we can write
v = r ;
where is the velocity potential. This potential can be found from the equation
of continuity (8.43):
= r2 =
280
2002 by CRC Press LLC
f
H0 0 ;
b
(8.44)
Redshift distortions
269
where we have used the present values of H and f . Thus, we can write
v=
H0 r(r 2 );
where f0 =b and r 2 is the integral operator that inverts the equation (8.44).
The radial velocity distortions can now be written as
s
;
(8.45)
(r) = S (r)
S is
@2
@r2
+ (rr) @r@ r
(x1 x2 ) = ( 12
1 2) = 1 2
S=1+
2:
(8.46)
= br r. The corre-
S S ( 12 ):
(8.47)
0x x
(s) = S
(s)
where
This is given by
S
where
@2
1 + @s2
1=1
0
+ s(s) @s@
r 2;
d ln 1 (s)
+ 2;
d ln s
(r) + 0 (r) = 4:
0 (s) =
and
Taylor and Valentine (1999) give also a clear review of different velocity distortion
effects and of their treatment.
8.3.2 Far-field approximation
The first studies of radial velocity distortions were done in the far-field approximation, where we can suppose that the density profile is a plane wave oriented at
an angle to the line-of-sight direction. Looking along the z -direction, we get
SP = 1 + @z@ 2 r
281
2:
270
S =1+
meaning that a redshift space Fourier mode is proportional to a real space Fourier
mode:
es
2
e :
(k) = 1 +
(k) (k)
P s (k) =
1 + 2 (k)2 P (k):
0:6
= M
b
(see Chapter 6 for the justification of the approximation) and its value is about 0.5
to 1.0. Thus, depending on the orientation, the redshift space spectral amplitudes
could be up to four times higher than real space spectrum; the velocity distortion
effect is substantial. Expanding the square, we can write
4
2
1
4
2
2
(k)=P (k) = 1 + 3 + 5 + 3 + 7 P2 () + 358 2P4 ();
where P2 () and P4 () are the Legendre polynomials that describe the quadrupole
Ps
(r) = 0 ( ) + 2 ( ) 2 ( ) + 4 ( ) 4 ( )
where
1
2
2
0 (r) = 1 + + (r);
3 5
4 + 4 2 (r) (r) ;
2 (r) =
3 7
5
7
8
2
4 (r) =
35 (r) + 2 (r) 2 (r)
and
Z r
= r33 (x)x2 dx;
Z0 r
5
(r) = 5
(x)x4 dx:
r
(r)
282
2002 by CRC Press LLC
Redshift distortions
271
=1
r=x x
= = cos =
= sin
= sk = js1 s2 j
and the transversal distance
= s? = s2 s2k ;
where si are the redshift distances for the two galaxies and s is the redshift space
separation of the pair.
Thus, the observed redshift space correlation function really depends on two
arguments,
; (in the far-field approximation).
The above derivation used the linear approximation to dynamics and thus describes only large-scale effects. There is as yet no theory to describe the effects
of nonlinear dynamics and this is approximated phenomenologically, supposing
that virial motions simply erase small-scale velocity correlations. This leads to
the convolution of the linear redshift correlation function by the one-dimensional
random pairwise velocity distribution of galaxies f v . In Fourier space it reduces
to the multiplication of the corresponding Fourier transforms.
Two model distributions have been used, the Gaussian velocity distribution
(r) = ( )
()
f (v ) =
p1 e
2
v2 =22
f (v) =
p1 e
2
283
2002 by CRC Press LLC
p2jvj=
kk of the wavevector)
272
Figure 8.9 The redshift space correlation function of the 2dF survey, shown as the gray
density map, and a model correlation function, shown with contour lines. The y -axis shows
the radial separation and the x-axis shows the transversal separation . (Reproduced,
with permission, from Peacock et al. 2001, Nature, 410, 169173. Macmillan Publishers
Ltd.)
fe(k) =
1
:
1 + 12 [k(k)]2
The velocity variance 2 here is the pairwise velocity variance that is two times
larger than the single-point velocity variance.
A model of the redshift space correlation function together with the results from
the 2dF redshift survey (Peacock et al. 2001), is shown in Fig. 8.9. The model uses
the estimated power spectrum from the APM survey that we shall discuss below,
: and the exponential pairwise small-scale distortion
and assumes a best-fit
km s 1 . Both the large-scale squashing of the correlation function
with
contours and the finger-of-God effect are predicted rather well by the model.
= 400
=04
273
E es (k)(es )? (k0 )
where
g(k; k0 ; k00 )
and
f (k; k0 ) =
= (2)3
Z
2 (r; k0 )
i(r)
(r; k0 ) ei(k
k0 r
(8.48)
(8.50)
nelmn = cln
? (; ) d3 s;
n(s)w(s)jl (kln s)Ylm
(8.51)
where cln are normalization constants, jl are spherical Bessel functions, kln are
discrete wavenumbers, and Ylm are spherical harmonics. We shall describe the
FourierBessel expansion in more detail in Chapter 9.
The pixelization (8.51) above gives pixels in the wavenumber space of spherical harmonics. Expanding the velocity, the data vector Dlmn is defined as the
normalized difference between the amplitudes (8.51) and similar harmonics for
the mean density (selection function). This leads to the relation
D = ( + V)de;
where e are the Fourier modes in real space and the matrices and describe
the mixing of modes due to the geometry of the survey (incomplete sky coverage,
for example) and the radial distortion, respectively. The choice of spherical coordinates allows us to separate the matrices and into the angular and radial
parts, writing
D = SW ( + V) de:
285
274
V
W
C = SW ( + V) P ( + V) W S + RW
where P is the diagonal prior power spectrum matrix and R is the radial part of
the noise matrix. The detailed expressions for these matrices are rather complex;
we refer the reader to the paper by Tadros et al. (1999).
In addition to the FourierBessel expansion there exist other possibilities. Hamis constant, the
ilton and Culhane (1996) note that when the selection term
radial distortion operator is scale-free and commutes with the operator @=@ r.
Then one can choose a complete set of commuting Hermitian operators:
(r)
3
+ =
@ ln r 2
ln
3
+ ; L2 ; Lz :
@ ln k 2
The last two (rotation) operators give the usual spherical harmonics and the first
operator logarithmic waves either in real space or in Fourier space. The orthonormal eigenfunctions for this set are
Z!lm (r) =
p1 e
2
(3=2+i!) ln r Y (br);
lm
where Ylm are spherical harmonics. When the observed number density is expanded in this basis, the redshift space Fourier amplitudes will be proportional to
in (8.52) will be diagonal, simthose for real space. It means that the matrix
plifying the calculation of the covariance matrix. Because the scales where we
measure power spectra usually extend over several decades, logarithmic pixelization in scale is better in this respect, too.
As an example of the application of the machinery described above we show
the real space spectrum obtained for the PSCz survey by Hamilton, Tegmark, and
Padmanabhan (2000). They expanded the galaxy density in logarithmic spherical waves, applied the KarhunenLo`ewe signal-to-noise transform, and used the
quadratic compression. This spectrum has also been corrected for linear redshift
distortions by allowing a scale-dependent . The resulting power spectrum for
large scales is shown in Fig. 8.10. As we see, the errors of the spectrum are especially large at large scales (a bit smaller for the correlated power spectrum). There
are interesting details in the spectrum at large scales but deeper surveys are clearly
needed to establish their reality. The model power spectrum shown describes observations rather well, except for the large wavenumber region, where nonlinear
dynamics may change the results.
286
2002 by CRC Press LLC
275
105
PSCz hi-b 0.6 Jy
12504 galaxies
104
103
CDM
102
.01
.02
.05
.1
.2
.5
Wavenumber k (hMpc1)
Figure 8.10 The real space power spectrum for galaxies from the PSCz redshift survey.
The solid line is the correlated power spectrum and the shaded region its rms error. Points
with rms error bars show the decorrelated power spectrum and the dashed line gives a
model power spectrum. (Reproduced, with permission, from Hamilton, Tegmark, and Padmanabhan 2000, Mon. Not. R. Astr. Soc., 317, L23L27. Blackwell Science Ltd.)
0x x
(x x )
=x x
The second angle is
, the angle between the ray bisecting the angle between 1
and 2 , and the direction of the far-away side 1
2 of the triangle. This angle
287
2002 by CRC Press LLC
276
cos
= cos (r1 r r2) :
s (r; ) = c00 0(0) + c02 2(0) + c04 4(0) + c11 1(1) + c13 3(1) + c20 0(2) + c22 2(2) ;
where
1
l(n) (r) = 2
2
k2 n P (k)jl (kr) dk
()
()
= 1 + 23 + 51 2 cos2 sin2 ;
4
4
2
1
4
2
2
2
2
2
c02 =
3 + 7 cos2P2 (cos
) 3 7 + 7 sin sin ;
4 2 sin2 P2 (cos
) 1 2 4 3 sin2 sin2 :
8
c04 = 2 P4 (cos
)
35
21
5 21 7
The functions Pn (x) above are Lagrange polynomials.
c00
i = ri (xi );
c11
c13
c20
=
=
c22
Matsubara, Szalay, and Landy (2000) used these formulae to find the redshift
space covariance matrix for the Las Campanas redshift survey, which has a rather
288
2002 by CRC Press LLC
277
complex geometry. They neglected the last four shape coefficients, integrated the
correlation function numerically for nearby pixels,
Sij =
Vi Vj Vi Vj
s (x1 ; x2 ) d3 x1 d3 x2 ;
2. Second, the data should be pixelized into pixels substantially smaller than
L= .
10
3. The pixelized data should be compressed using the KarhunenLo`ewe transform. Although this compression is not too strong, it conserves phase information and the compressed data can be used both to estimate velocity distortions
and to analyze the assumptions made in the analysis; these show up clearly in
the error distribution.
4. Quadratic compression should be used to get the band-powers; the final maximum likelihood analysis is most efficiently done using the band-powers as the
final data vector.
As an example of such a combined approach we demonstrate in Fig. 8.11 the
latest estimate of the real space power spectrum of the PSCz redshift survey over
the whole accessible wavenumber range by Hamilton and Tegmark (2000). The
:
h Mpc 1 to
wavenumber range for this power spectrum extends from k
k
h Mpc 1 , almost 4.5 decades.
The linear part of the spectrum has been obtained by the machinery described
above and is the same as that shown in Fig. 8.10. As we noted above, it is not
particularly usable for small scales; the density distribution is not Gaussian any
more and the number of pixels will be prohibitively large. Here the nonlinear part
has been obtained using a version of the direct method. Because the redshift space
spectrum is anisotropic, it was expanded into spherical harmonics. The harmonics
of the band-powers were defined as
= 0 0137
= 316
Pls
(k; ) =
d3 k 0
W (k; k0 )(2l + 1)Pl ()P s (k0 )
(2)3 ;
289
(8.53)
278
105
= 0.62, m = 0.38
bh2 = 0.020, h = 0.63
n = 0.905, b = 1.24
103
no
102
nl
in
ea
r
10
lin
linear
nonlinear
ea
104
101
102
.01 .03
.1 .3
1
3
10 30
Wavenumber k (h Mpc1)
100 300
Figure 8.11 The real space power spectrum for galaxies from the PSCz redshift survey. The
spectrum is a combination of the linear calculations and the direct estimator in the nonlinear regime. The solid line shows the correlated power spectrum and the shaded region
shows its rms error. Points with rms error bars show the decorrelated power spectrum in
the linear wavenumber region and the correlated power spectrum in the nonlinear region.
The dashed line gives the power spectrum for a CDM model with the parameters shown in
the figure. (Reproduced, with permission, from Hamilton and Tegmark, Mon. Not. R. Astr.
Soc., in press. Blackwell Science Ltd.)
)
(k
()
Pls (k; ) =
where Wl
il
(k0 )2 dk0 :
jl (kr)W (k; k0 )
(2)3
0
290
2002 by CRC Press LLC
279
( )
ihRRi
1 + s (r; ) = hDD
hDRi2 ;
where the quantities in angle brackets are data and random reference pair counts,
the latter done here as integrals over the catalog volume. For a direct method the
band-power windows can be chosen by hand; Hamilton and Tegmark chose
0 n
0
2
k
W k; k0
e (k =k) :
(0 ) = 0
= 72
ln
1 12
( )= ( =0
= )
where kk and k? are the radial and transverse wave vectors. Hamilton and Tegmark argue that the estimate (8.54), although unbiased, is rather noisy and the use
of the redshift information is equivalent to improving the spectral resolution by
R= H0 times, where R is the depth of the survey and 2 is the typical onedimensional velocity variance. For the PSCz survey this ratio is about 16.
Redshift distortions can be quantified by defining the correction function
Ps
f (k) =
(k) :
P (k )
As their analysis showed that the standard pairwise distribution functions did not
describe this ratio well enough, Tegmark and Hamilton (2000) approximated the
correction function by a finite sum over even harmonics
f (k) =
lmax
(k)
X
() ()
fl k Pl :
l=0
The real space band-powers can be obtained then as a sum of the redshift space
band-power harmonics, weighted by the correction function. The maximum hark1=2 to resolve
monics was chosen as the nearest even integer to lmax k
radial scales corresponding to the standard one-dimensional rms velocity error
=H0 h 1 Mpc.
Because the density distribution at small wavelengths is not Gaussian, we cannot use the Fisher matrix to estimate the variances of the band-power estimates.
( ) = 16
=3
291
2002 by CRC Press LLC
280
Hamilton and Tegmark (2000) found the variances using the fluctuations of the
estimates in the survey. The catalog was divided into a number of subvolumes,
and the fluctuation of a band-power estimate for a subvolume i was defined as
@ Pb
Pbi = 12 wi @w
;
i
hPb2 i =
ij
PbiPbj :
(8.55)
The problem with this method is that the sum of fluctuations over all subvolumes
vanishes, leading to a zero variance, and we have to restrict the number of pairs in
the sum (8.55). This is done by including into the sum (8.55) pairs of subvolumes
with growing separation and canceling summation when the sum has attained its
maximum value (Hamilton 1993b).
The resulting power spectrum is pretty close to a power law over a large wavenumber region. As a check of the methods used, the linear and nonlinear parts
of the spectrum match well. The full line shows the correlated estimate of the
power spectrum and the shaded region shows its rms error for linear scales. The
error bars show the decorrelated spectrum for linear scales; because there is no direct analog of that for nonlinear scales, the error bars there describe the correlated
spectrum. The model spectrum shown in the figure by a dashed line does not fit the
observations well; the easiest way out is to postulate scale-dependent biasing. The
power spectrum has two regions where errors are large. The high wavenumber region describes scales that are too close to the sizes of galaxies; a meaningful limit
to the galaxy power spectra would be around wavenumbers k 30100 h Mpc 1
( 30100 h 1 kpc). The large errors in the small wavenumber limit are due to
insufficient depth of the catalog and also show that the Fisher matrix approximation for rms errors does not work well in that region. The solution is to use there a
full-blown maximum likelihood method that gets confidence regions by integrating over the parameter (band-power) space. And, as before, we see that there is
not yet enough evidence for a turnover in the power spectrum. It must be rather
close, because the CMB spectrum measurements show the mean power spectrum
h 1 Mpc (k : h Mpc 1 ).
of opposite slope P k k at scales near ;
()
1 000
0 003
8.6 Bispectrum
The bispectrum is identically zero for Gaussian fields. Thus a nonzero amplitude
of the observed bispectrum will be a clear indication of deviations from Gaussianity. However, because many different mechanisms may generate such deviations,
it is rather difficult to entangle different contributions to the bispectrum.
292
2002 by CRC Press LLC
Bispectrum
281
(8.57)
Perturbation theory for gravitational dynamics (Fry 1984) predicts a similar expression in the second-order approximation (in first order the bispectrum remains
zero). It gives, however, different coefficients for every term in (8.57):
where we have used a notation similar to that used for correlation functions, P1
P 1 , etc. The coefficients Qij are determined by the geometry of the triangle
f g:
(k )
123
ki
kj
+ kkji + (1
)cos2 ij ;
where ij is the angle between the two sides of the triangle with lengths ki and
kj . The coefficient describes the cosmological model and varies only slightly
= . Velocity distortions and bias also retain the structure of this
around
expression, changing only the coefficients Qij in a complicated way (Verde et al.
1998, Scoccimarro 2000). This leads to the definition of the reduced bispectrum
=37
Q123 =
B123
:
P1 P2 + P2 P3 + P1 P3
(8.58)
While the power spectrum and the bispectrum vary strongly with a typical
wavenumber, for a power-law spectrum the reduced bispectrum depends mostly
on the shape on the triangle formed by the vectors 1 ; 2 ; 3 and its amplitude
does not vary much. We show the expected bispectrum for different physical models in Fig. 8.12. It is higher for elongated triangles, depends on the assumed spectral index, depends little on the cosmological model, and velocity distortions do
not change it much.
k k k
293
2002 by CRC Press LLC
282
Figure 8.12 The dependence of Q on the shape of the triangle. The power spectrum
P (k ) k 1:4 . The triangles are described by the side ratio k1 =k2 = 2 and by the angle
between these sides (the horizontal axis). The solid curve shows the result in real space, the
short-dashed curve in redshift space for
M = 0:3, and the long-dashed curve in redshift
space for
M = 1. The dotted curve stands for the case
M = 0. The dot-dashed curve
shows the effect of a (small) local bias. (Reproduced, with permission, from Scoccimarro
et al. 2001, Astrophys. J., 546, 652664. AAS.)
b
gal = b1 + 2 2 :
Qgal =
Q
b1
+ bb22 :
1
The main effect is the change of the overall amplitude of the bispectrum. The
dot-dashed curve in Fig. 8.12 describes the case of a rather mild (anti)bias, b1
: ; b2
:.
As the galaxy bias b is usually difficult to estimate (velocity distortions give the
f0 =b), determination of the bispectrum can give independent
combination
evidence on b.
08 = 03
294
2002 by CRC Press LLC
Bispectrum
283
(k )
12 23 + 23 31 + 31 12 + 12 23
D = D (x1 x2 ) for brevity. This is simiwhere we have used the notation 12
1 2 3
123 +
lar to the formula for the two-point correlation (8.9) above; here there are more
discreteness correction terms.
We can use a similar derivation as in Section 8.2 to get the expectation value
E fF1 F2 F3 g. This, however, leads to complicated expressions. Matarrese, Verde,
and Heavens (1997) derived a simpler formula, using the approximation of narrow
window functions, that allows us to get the power spectra out of the integrals. In
our notation of Section 8.2 it gives
E fF1 F2 F3 g
B123
+
(x)
3 (x) d3 x + (P + P + P )
1
2
3
3 (x)
3
2 (x) d x;
V n
3 (x)
(x)
V n
d3 x +
(8.59)
(x)
where
is the weight function and n
is the local mean density of the
survey. The estimator for the power spectrum is in the same approximation
E fjF1 j2 g = P1
2 (x) d3 x +
2 (x)
(8.60)
d3 x:
V
V nx
Scoccimarro (2000) has shown that the optimal weight function x for the bispectrum is the same FKP function (8.19) that was found to be optimal for estimation of the power spectrum. Normalizing the weights for the power spectrum
by
Z
( )
()
2 (x) d3 x = 1
Pb = hjF j2 i I21 ;
where the average is over a shell in the wavenumber space, and
Bb123 =
30
295
2002 by CRC Press LLC
30
284
Figure 8.13 PSCz survey bispectrum for triangles of the side ratio k2 =k1 = 0:4 0:6,
parameterized by the angle (horizontal axis). The solid curve is the result for all matter
in a numerical CDM model, and the dashed curve shows the model result for galaxies,
obtained using the best-fit bias parameters (see text). Symbols show the observational results for triangles of different size k1 : filled triangles for 0.200.24 h Mpc 1 ; filled squares
for 0.240.28 h Mpc 1 ; filled circles for 0.280.32 h Mpc 1 ; open circles for 0.320.36
h Mpc 1 ; and open squares for 0.360.42 h Mpc 1 . (Reproduced, with permission, from
Feldman et al. 2001, Phys. Rev. Let., 86, 14341437. Copyright 2001 by the American
Physical Society.)
This result was obtained first by Matarrese, Verde, and Heavens (1997). Here
the assumption that the random density field is isotropic allows us to average
over all orientations of the triangles of the same size and shape as f g in the
wavenumber space. We have used the notation
Z
N
INM
d3 x
M
V n
for the normalization integrals above. The reduced bispectrum estimator is then
123
Qb123 =
(x)
(x)
Bb123
:
Pb1 Pc2 + Pb2 Pb3 + Pb1 Pb3
(8.61)
While the estimators for the power spectrum and for the bispectrum are unbiased, at least in the approximations used above, the estimator for the reduced
bispectrum is biased. Moreover, its one-point distribution is skewed, as found by
Scoccimarro (2000), so care is needed when fitting observed estimators to models.
296
2002 by CRC Press LLC
Low-dimensional samples
285
The bispectrum estimates are rather noisy. The reason for that is simple as
explained at the beginning of this chapter, there is a finite number of coherence
cells in the volume of the wavenumber space that we can study. The estimates of
the power spectra are improved by assuming isotropic spectra and averaging individual estimates over k -space shells. The parameter space for the bispectrum is
much larger, we can average only over identical triangles of different orientation,
and, consequently, the number of coherence cells that can be used to get a single
estimate of the bispectrum is small.
In conclusion, we show in Fig. 8.13 the bispectrum of the PSCz survey, obtained by Feldman et al. (2001). Different symbols show different triangle configurations and the lines show the model results (solid line means unbiased, dashed
line means the best bias fit). We see that, although the scatter is large, the reduced
bispectrum is definitely larger than zero and follows approximately theoretical
predictions. The estimate for the linear bias =b : is close to the PCSz survey
bias b : found from velocity distortions, and the fit also gives a nonlinear
:.
bias term b2 =b21
0 84
12
05
286
(r) r
Z 1
(br) =
(8.62)
(y)y2 (y) dy
o
where (y) is the real density contrast,R y = jyj is the radial distance, and (y ) is
the selection function, normalized by (y )y 2 dy = 1.
depending by isotropy only on the angle between the two directions. Using the
expression (8.62), we get
Z Z
w ( ) =
(8.63)
Two-dimensional surveys are usually deep, with the depth L much larger than
the correlation radius r0 . This means that we can use the small-angle approximation:
r2 = jy1 y2 j2 = x2 + y2 2 ;
where x = y1 y2 and y = (y1 + y2 )=2. Changing to the variables x and y in the
integral (8.63) we get
w ( ) =
y4 2 (y) dy
p
x2 + y2 2 dx:
(8.64)
This formula is known as Limbers equation. The integration limits are formal,
reflecting the assumption that y is different from zero only in a limited interval.
This formula predicts that the angular correlation function scales with the sample depth D. Let us define this depth by y > D . The normalization
condition tells us that the overall amplitude of D 3 . If the distribution of
galaxies is homogeneous on average, the main part of the integral comes from the
regions y D and the correlation function can be written as
()
w()
1 W (D):
This scaling relation has been checked in practice (see a review in Peebles 1980,
x 57). It confirms the theory and also provides one of the arguments against the
fractal nature of galaxy distribution at large scales. As an example, Fig. 8.14
shows the scaling of the angular correlation function of the APM galaxy survey (Baugh 1996). The different correlation functions were obtained by selecting
galaxies by their apparent magnitude the larger the apparent magnitude, the
larger the mean depth of the survey. As expected, the shape of the angular correlation function remains the same, but it is shifted (by the same amount) both in
the log-amplitude and in the log-argument.
298
2002 by CRC Press LLC
Low-dimensional samples
287
-1
-2
-3
-1
-0.5
0.5
Figure 8.14 Scaling of the angular correlation function of the APM galaxy survey. The
different functions correspond to the magnitude intervals of 17 bJ 18, 17 bJ 19,
and 17 bJ 20, counting from above. For clarity, the errors of the middle curve have
been omitted. (Reproduced, with permission, from Baugh 1996, Mon. Not. R. Astr. Soc.,
280, 267275. Blackwell Science Ltd.)
As photometric surveys are usually deep enough to feel the curvature effects,
the relativistic version of Limbers equation is commonly used. Let us recall the
definition of the volume element from Chapter 2
instead of (8.63)
w(!1 ; !2 ; ) = R06
Z Z
(8.65)
(the Mattig comoving distance), we can transform the relativistic Limber formula
299
2002 by CRC Press LLC
288
w() =
2 (y)
y4
dy
Ck (y)
p
x2 + y2 2 dx:
()
( )= 1
(8.66)
Ck (y)
R0
it is the
=0
= 1
(y)
y2
dy = 1:
Ck (y)
The formula (8.66) describes the situation when the three-dimensional separation is measured in comoving coordinates (the units of R0 ! are h 1 Mpc). Many
papers use a slightly different form of Limbers equation, based on a different
form for the space-time metric and using physical separations:
w() =
where
y4
2 (y)a6 (y) 2 dy
F (y)
(y) is normalized by
Z
(y)a3 (y)
()
p
x2 =F 2 (y) + y2 2 dx;
y2
dy = 1
F (y)
()
()
Low-dimensional samples
289
=1
=0
() ( )
(r; z ) (1 + z ) 3 :
The factor
in the exponent arises from the normalization of the density contrast
z 3 . This case is called stable clustering and that
by the mean density
assumption has been used frequently to describe the evolution of correlations.
Elaborate machinery has been developed to approximate the evolution of the
correlation function (and the power spectrum) in the intermediate regime between the two asymptotes. This is known as the HKLM procedure after its authors
(Hamilton et al. 1991). The idea of this procedure is to represent the nonlinear correlation function as a universal function of the linear correlation amplitudes for
rescaled arguments. Peacock and Dodds (1996) extended this procedure to predict
the nonlinear power spectrum on the basis of the linear power spectrum. As the
HKLM procedure is analytical, it is used rather frequently to predict nonlinear
power spectra for large parameter spaces of cosmological models, where direct
N -body modeling would not be feasible. The HKLM procedure is explained in
detail in Peacocks book (1999).
The selection function y in Limbers equation (8.66) is usually estimated by
noting that
(1 + )
()
(y)
y2
dN
dy =
dz;
Ck (y)
dz
(8.67)
50
0 15
301
2002 by CRC Press LLC
290
(r; z ) =
r
r0
f (z );
as we did above when discussing evolution effects. Then the angular correlation
function is
w A1
;
(8.68)
( )=
where the coefficient A depends on the galaxy redshift distribution, on the cosmological parameters, and on the assumption about the evolution of the spatial
correlation function. Finding the amplitude and the exponent of the observed angular correlation function, we can solve (8.68) to find
and r0 .
As the exponent of the correlation function does not depend in this approximation on any unknown parameters, it is a useful probe of the shape of the spatial
correlation function in the past (at large redshifts). Giavalisco et al. (1998) applied this technique to a sample of distant Lyman-break galaxies at z and
found that their spatial correlation function has about the same slope as that of the
present-time galaxies. As their sample spans a relatively small redshift interval,
they also could determine the correlation length without making any assumptions
about the evolution of the correlation function. Their results show that the correlation function does not change appreciably with redshift.
(r) = 4
w() = 2
()
P3 (k)
2 (y)y4 dy
sin(kr) k2 dk ;
kr (2)3
P3 (k)J0 (ky)
k dk
(2)2 ;
(8.69)
()
( )
!() = 2
P2 (K )J0 (K)
302
K dK
(2)2 ;
Low-dimensional samples
291
P2 (K ) =
( )
P3
K 2 2
(y)y dy:
y
(8.70)
()
P2 (K ) 3 p(K=D);
D
()
w ( ) =
(; k) is given by
Z 1
dN 2
1
Gw (; k) =
2R0 0 dz (1 + z)
0 ky (z )
(8.71)
d!
dz
1
dz:
The second factor in the integrand describes the evolution of the spatial correlaR0 Sk ! z and the curvature radius
tion function. The functions ! z , y z
R0 depend on the parameters of the cosmological model. The factor Ck y is canceled by a similar factor from the derivative dy=dz . The power spectrum of galaxy
distribution can be found now from (8.71) by deconvolving the angular correlation
function. The kernel Gw ; k is rather wide and oscillating, so deconvolution is
not easy.
The formula for the two-dimensional spectrum reduces in this approach to
( ) ( )=
( ( ))
()
( )
KP2 (K ) =
where the kernel GP
(y) is
GP (y) =
()
2
dz
dN 2
(1
+
z ) Ck (y)
dz
dy
(8.72)
()
303
2002 by CRC Press LLC
292
Figure 8.15 Power spectrum from the APM galaxy survey. Filled circles show the maximum likelihood estimates, and open circles show the spectrum obtained by deconvolution
of the two-dimensional power spectrum. Lines show model predictions. (Reproduced, with
permission, from Efstathiou and Moody 2001, Mon. Not. R. Astr. Soc., 325, 16031615.
Blackwell Science Ltd.)
Baugh and Efstathiou (1993, 1994) determined the scatter of the power spectrum estimates by breaking the survey into four distinct regions. Both methods
gave a similar result that was long considered the most reliable power spectrum
estimate. The size of the APM survey was much larger than that of any redshift
surveys available, and this estimate does not depend on the treatment of redshift
distortions. In a recent paper, however, Efstathiou and Moody (2001) re-estimated
the APM power spectrum by applying the maximum likelihood method to the
APM pixel map. The results are compared in Fig. 8.15. The maximum-likelihood
results are shown by the filled circles and the results of the previous deconvolution are shown by the open circles. While the results agree rather well, the clear
maximum around k : h Mpc 1 that is seen in the sequence of open circles is
not confirmed by the ML method (the ML point at k : h Mpc 1 is outside
the graph). This shows once more that direct methods do not give reliable results
for scales close to the size of the survey volume.
0 02
0 013
304
2002 by CRC Press LLC
Low-dimensional samples
293
f (x) =
()
(8.73)
( )
where f x and K x; are known, we can rescale all these functions first to
have a unit integral:
Z
f (x) dx =
g() d =
K (x; ) dx = 1:
()
()
( )= ( )
g() =
(8.74)
( )
where L jx is the conditional probability density for , given x. This density
can be obtained from Bayes theorem:
L(jx) =
g()K (xj)
:
f (x)
(8.75)
()
( )
Lucy (1974) showed that we can now proceed by iterations, choosing g first
to be a reasonable smooth function (Baugh and Efstathiou chose a power law for
P3 k ). Denoting the first choice g1 , we find f1 x from (8.73) and L1 jx
from (8.75). Using that in (8.74) we get the next approximation g2 . This process leads to the rule
()
()
()
()
()
()
f (x)
K (xj) dx;
fi (x)
where fi x and gi are connected by (8.73). Lucy (1974) analyzes the stopping criterion and shows that the best results are usually obtained after a few
iterations; further iterations amplify small-scale features in g that are caused
by (frequently random) fluctuations in f x .
()
()
120
305
2002 by CRC Press LLC
294
this paper Peebles uses different Fourier conventions and also treats point source
samples.
For a wide-angle survey we can expand the density contrast as
(br) =
1 X
l
X
l=0 m= l
am
l Ylm (br);
( )
where b is the unit vector of the direction in space and Ylm ; are spherical
harmonics. Spherical harmonics (see Abramowicz and Stegun 1965, Press et al.
1992) are defined as
Ylm (; ) =
()
(8.76)
sin
= cos
( ) ( )
w()
E am 2
jlj
1 X
l
X
2
= 41
E jam
l j Pl (cos );
l=0 m= l
(8.77)
= 2
(8.78)
Z 1
()
2
The functions Pl x are the Legendre polynomials; note that jam
l j does not depend on m.
The exact Limbers equation does not need spherical harmonics; the nonrelativistic version reads
w ( ) =
1Z 1
r2 = s2 + t2
2st cos :
The selection function is normalized as above, and relativistic generalization proceeds in the same fashion.
The relation between the power spectra can be written as (Peacock 1999):
E am 2
jlj =
306
2002 by CRC Press LLC
k2 dk
(2)3 ;
Low-dimensional samples
295
(k) is
1
jl (ky)(y)y2 dy;
Gl (k) = 4
()
40
20
130
( ) = (0 0 )
P1 (k) =
(8.79)
where the last integral comes from the change of integration variable to
y 2 = k2 + r2 .
dP1 (k)
dk
= 2 kP3 (k):
307
296
Figure 8.16 Distribution of galaxies in the Broadhurst et al. (1990) pencil-beam survey.
The upper panel shows the binned distribution along the beam; dashed lines mark the
130 h 1 Mpc periodicity. The lower panel shows the one-dimensional power spectrum of
the survey. (Reproduced, with permission, from Yoshida et al. 2001, Mon. Not. R. Astr. Soc.,
325, 803816. Blackwell Science Ltd.)
()
P1 (k) =
()
fz j2 (kz k )jW
f j2 (kx ; ky )P3 (k ) d k ;
jW
r
(2)3
fz k and W
f kx ; ky are the Fourier transforms of the selection functions
where W
along the beam and for the transverse directions, respectively. If we approximate
308
2002 by CRC Press LLC
Low-dimensional samples
297
p
1 Z 1 jW
f j2 ( y 2 k 2 )P3 (y )y dy;
2 k
f (k ) = 2J1 (kR)=(kR) (J1 (x) is the Bessel function of order 1). The
where W
spectral window is a disk with the width kr 1=R The thickness of the window function along the beam, which was ignored in the approximation above, is
kz 1=L. For the Broadhurst survey the ratio of these dimensions is
kr =kz 100.
P1 (k) =
A pencil-beam is really a cone; Kaiser and Peacock (1991) derive the exact
formulae for that case, but the results do not differ much.
Thus, the spectral resolution of a pencil-beam survey is determined by its transverse dimensions and not by its depth. Based on that, most of the papers devoted to
the analysis of the Broadhurst et al. (1990) data discarded the result as a statistical
fluke, caused by the high-frequency modes. The case is open, however, because
the periodogram analysis of the distribution of distances between distinct clumps
of galaxies showed that it is too regular to be explained by random models (Dekel
et al. 1992). Extensive modeling of pencil-beams by numerical simulations also
has been unable to reproduce the regularity at that level (see Yoshida et al. 2000).
The case of the two-dimensional slices is similar to that above. The power
spectrum for an infinitely thin slice is
The three-dimensional windows for a slab (slice) resemble long needles and
the spectral resolution is again limited by the smallest dimension of the survey region. Because of that, the power spectra of low-dimensional samples are severely
distorted by the leakage from the high-wavenumber regions. The correlation functions of pencil-beams and slices differ from the spatial correlation functions only
by greater large-scale noise. This illustrates clearly the basic difference between
the power spectrum and the correlation function while the power spectrum
separates different scales (as well as possible), correlation functions mix contributions over all scales.
However, if we relax the assumptions about isotropy, then the pencil-beams
and slices can help tremendously in measuring the anisotropic features of galaxy
distributions.
309
2002 by CRC Press LLC
CHAPTER 9
Cosmography
9.1 Introduction
The basic statistical model used for the description of the spatial distribution of
galaxies is the inhomogeneous Poisson point process, where the intensity function
is determined by a realization of a Gaussian random field. The goal of cosmography (cosmic cartography) is to study this realization. The motivation is, first, to
see and understand the universe around us. Second, the better we map our local
neighborhood, the better we are able to answer the questions about its statistical
properties and its past.
As usual, we start with frequentist methods. The simplest methods of density
estimation are to use counts-in-cells or kernel estimators on the observed survey
data. As the density of galaxies varies greatly, we could use adaptive kernels (Silverman 1986) or wavelets. Adaptive density estimation methods are usually used
for discriminating structures (groups and clusters of galaxies) in redshift space
(see, e.g., Pisani 1996).
If we wish to find the local intensity in real space, we have to worry about velocity distortions, as the survey data is usually obtained for redshift space. We have
seen above that the differences could be of the same order as the densities itself, or
even much more recall the finger-of-God effects. While this distortion is better
to quantify analytically when we determine power spectra, for map-making the
first goal would be to get rid of these fingers. In our immediate neighborhood, for
example, rich clusters of galaxies distort severely the galaxy maps (see Fig. 8.4,
the Slice of the Universe, where the the stick-man really does not have a body;
it is the finger-of-God of the Coma cluster).
This subject has not raised much interest in recent years. The first attempt to
correct for these fingers was made by Gramann, Cen, and Gott (1994). They found
the cluster fingers by a clustering algorithm (usually called friends-of-friends in
cosmology), where the neighbor distance was scaled differently in the radial and
transversal directions. This scaling can be determined from the observed density
profiles of clusters of galaxies in redshift space; the scaling ratios usually are
about 610. After identifying the cluster fingers, they collapsed those to disks at
the median redshift of the cluster. They were not interested in cartography; a better
way was to compress them to spherical distributions, as was done later by Monaco
and Efstathiou (1999) and Monaco et al. (2000) for the PSCz reconstruction. Of
course, because masses and mean densities of galaxy clusters vary, the scaling
310
2002 by CRC Press LLC
300
Cosmography
ratios vary, too. The compression of cluster fingers really needs an application
of adaptive matching filters, as proposed for automatic identification of galaxy
clusters from photometric and redshift surveys (Kepner et al. 1999). The idea is
to fit parametric cluster models over the whole survey volume and analyze the
peaks of the resulting likelihood map. This demands serious computer resources
and is only now becoming feasible.
Another problem with direct density estimation is that of bias we cannot be
sure that the galaxy distribution traces the total distribution of matter. Fortunately,
as gravitational forces are universal for all types of objects, there are ways to map
the total density.
9.2 Potent method
Although the main distance measurement method by far is by galaxy redshifts,
there exist independent methods of distance determination. Such methods are
based on the distance ladder described in Chapter 1, and they are very laborintensive and rather inexact, the best instrumental errors being about 20%. Thus,
these distance estimators cannot be used directly to estimate densities of largescale structures. However, the existence of independent distance estimators allows
us to delineate the peculiar velocity fields. The first attempt was made by LyndenBell et al. (1988). They estimated the distances and peculiar velocities for 400
elliptical galaxies and discovered that there exists a mean flow toward a region
behind the galactic plane that they called the Great Attractor. This region had
not been detected before and the velocity fields were useful in describing the local
density distribution.
This result motivated Bertschinger and Dekel (1989) to propose a method to
recover the full velocity field. They used the fact that in the linear approximation
of structure evolution the velocity field is irrotational,
v(x) = r (x);
(x)
(9.1)
(x)
where
is the velocity potential. Thus, the observed radial velocity vr
contains the same amount of information as the velocity potential, and the potential can be recovered by integrating the radial velocity along radial rays from the
observer,
Z r
r; ;
vr r0 ; ; dr0 :
(9.2)
)=
Once we have the potential, we can find the full velocity field by (9.1). In the
linear approximation, the continuity equation gives
1r v
(9.3)
(f 0M:6 is the nondimensional density growth rate), so we can find the local
density distribution. The important point here is that in (9.3) describes the full
311
2002 by CRC Press LLC
Potent method
301
density distribution, dark matter plus galaxies, and thus it is free of assumptions
about biasing.
The practical realization of the method, called POTENT, was proposed by
Dekel and Bertschinger (1990). Although the method is simple in principle, several sources of possible error were found during its application in later years. The
second version of POTENT (Dekel et al. 1999) has been designed to eliminate
these errors. The steps of POTENT are as follows:
1. Correction of observed peculiar velocities. As the independent distance determinations reduce to the estimation of the absolute luminosities of galaxies,
these distances suffer from the Malmquist bias, and must be corrected. The
types of Malmquist bias depend on the distance determination method. A thorough review of different types of Malmquist bias for peculiar velocity catalogs
and correction methods is given in Strauss and Willick (1995). Correction of
distances implies correction of peculiar velocities.
2. Smoothing of the observed radial velocity field. This has to be done for two
reasons first, the formulae written above are based on the linear approximation of the evolution of structure that is applicable only for comparatively large
scales, at about a few tens of Mpc and larger. Second, the peculiar velocity
surveys are sparse. The largest survey done thus far is the MARK III survey
(Willick et al. 1997), which includes about 3,300 galaxies in a volume of a radius of about 90 h 1 Mpc. So, in order to be able to integrate along radial rays
we have to interpolate the radial velocity field first over the survey volume. The
smoothing kernel chosen is a spherical Gaussian, and due to sparsity of data
the width of the window RG is typically chosen at about 1012 h 1 Mpc.
Because the radial directions in neighboring regions differ, the interpolation
has to use tensor weights and has to assume a velocity model. Typical models
;j
used are a constant flow near a point
i j RG or a linear
shear field
i , where has to be a symmetric matrix
to ensure that the flow is irrotational. The combination of tensor weights with
a spherical window introduces a bias called the tensor window bias. Using
higher-order models can reduce that bias.
v(x) = B + T(x x )
v(x) = B x x
T
As the mean radial velocity changes through the window, nonuniform sampling gives rise to a bias that is called the sampling-gradient bias. It can
be reduced by using volume-weighting recipes for interpolation instead of the
straightforward galaxy weighting.
The velocities have also to be weighted to take account of the random distance
measurement errors.
3. Integrating the interpolated velocity field to recover the velocity potential and
the three-dimensional velocity field.
4. Estimating the density field. The formula (9.3) is only true for small values of
density contrast and we can improve on that.
312
2002 by CRC Press LLC
302
Cosmography
The only assumption we have used in recovering the velocity field thus far is
the existence of the velocity potential. Recall that in the Zeldovich approximation the flow of matter stays irrotational until the first crossings, up to infinite
density. Thus, in this step methods are used that also work in mildly nonlinear
regimes to improve the dynamic range of the prediction.
There are several higher-order approximations that describe the structure evolution. The POTENT choice is the Eulerian version of the Zeldovich approximation (Nusser et al. 1991). The Zeldovich approximation predicts presenttime Eulerian positions of mass elements (objects) as a function of their Lagrangian coordinates:
1 v(q);
x(q) = q + f (
)
(9.4)
where the function f is the dimensionless growth rate of perturbations. Assuming that there have been no flow crossings, the mapping (9.4) is unique and can
be inverted:
f 1
:
(9.5)
q(x) = x
v [q(x)]
x d3 x = q d3 q = d3 q
(x) =
v
@x
@
f 1 :
(9.6)
This (rather crude) approximation works better than the higher-order expressions. For POTENT use, the Jacobian in (9.6) is modified by changing slightly
the coefficients of powers of f (Dekel et al. 1999).
5. As the errors of this complex analysis are difficult to estimate, the last step
includes evaluation of the errors, using the POTENT method on mock N -body
catalogs to find the influence of observational effects.
The results of the POTENT analysis provide an independent measure of the
total mass density and thus help to clarify the nature of the galaxy bias. Similarly,
POTENT gives also a direct way to estimate M (recall that the dependence of f
on is small; see Chapter 6).
The density and velocity maps of our local neighborhood obtained by POTENT
h 1 Mpc)
are shown in Fig. 9.1. Due to the wide filter (the rms width is RG
the map is rather smooth. The peculiar velocity field shows a smooth flow over
the region converging at the Great Attractor. The relief map in the lower panel
shows the main features of the local density distribution the Great Attractor
(GA); the Great Wall (GW) that was first seen in the Slice of the Universe; the
PerseusPisces supercluster (PP) and its extension, the Southern Wall (SW); the
Local Group of galaxies (LG) in the center and the Sculptor Void (Void). The
Local Supercluster itself is the low shoulder to the left of the Local Group.
= 12
313
2002 by CRC Press LLC
Potent method
303
Figure 9.1 POTENT maps of the supergalactic plane from the Mark III peculiar velocity
catalog. The upper left panel shows the density contrast: the heavy contour corresponds
to
, darker areas have larger densities, and the contour spacing is
: . The
upper right panel shows the projections of the peculiar velocities on the supergalactic
plane. Distances are in units of h 1 Mpc. The lower panel shows the relief of the density
contrast. (Reproduced, with permission, from Dekel et al. 1999, Astrophys. J., 522, 138.
AAS.)
=0
= 02
This distribution is similar to the reconstruction of the local neighborhood obtained by different methods, showing that gravitation is really the main force behind the evolution of structure. Comparison of reconstructions on densitydensity
or velocityvelocity basis helps clarify further our picture of that evolution.
There are several reconstruction and velocity comparison methods similar to
POTENT; a review of these methods is given by Dekel (1997). These methods
can be roughly divided into two classes. The first class is represented by POTENT,
which attempts to reconstruct the velocity and density fields in real space. Comparison of the reconstructed density field with density fields found from redshift
bf M (b is the bias and
surveys allows us to determine the parameter
f M is the dimensionless density growth rate). These methods are called the
densitydensity (d-d) comparison methods.
= (
)
(
)
314
2002 by CRC Press LLC
304
Cosmography
The second class contains the methods that attempt to reconstruct the observed
velocity fields in redshift space and to compare them with the velocity fields found
from redshift surveys. The best known of these methods is the VELMOD maximum likelihood method (Willick et al. 1997), which reconstructs velocity fields
by maximizing the probability of observing the peculiar velocity indicators (apparent magnitude and velocity width) of the sample galaxies, given their redshifts.
Because in this case is estimated by comparing velocity fields, these methods
are called the velocityvelocity (v-v) comparison methods. Thus far, the value
of found by (d-d) methods is about twice the value found by the (v-v) methods; the reason for that is not yet known. The latest velocityvelocity analysis by
Silberman et al. (2001) gives the estimates of M that are close to the estimates
found from the analysis of the CMB data and of the deep redshift surveys.
d = Rs + n;
(9.7)
s = Fd
and demand that it be optimal in the sense of minimizing the variance of the
discrepancy
r = s bs;
where
is
s is the true value of the signal vector. The variance we have to minimize
hrry i = (s Fd)(sy dy Fy ) ;
where the angular brackets denote averages over realizations. As the data and
signal vectors also may describe the Fourier amplitudes, we have to use Hermitian conjugation, denoted by a dagger (in the case of real matrices that will
315
2002 by CRC Press LLC
Wiener filtering
305
reduce to transposition). Adding to both sides of the above expression the term
h y ih y i 1 h y i (completing the square) we get
sd dd
ds
hrry i = F hsdy ihddy i 1 hddy i F hsdy ihddy i 1 y
hsdy ihddy i 1 hdsy i + hssyi:
F gives
F = hsdyihddyi 1:
(9.8)
DS
where , , and
are the data, signal, and noise covariance matrices, respectively. This derivation assumes that data and noise are not correlated. The first
term in (9.8) reduces to
s = SRy(RSRy + N) 1d:
(9.9)
n = Rn
s = S(S + N0) 1R 1d
(9.10)
(where 0 is the new noise covariance matrix) and states that the estimator is
obtained by inverting the initial expression (9.7) and weighting the obtained result by the ratio of signal to signal-plus-noise. The mean square residual can be
transformed to
y
y
1
y 1;
h yi
(9.11)
rr = SR (RSR + N) N(R )
hrry i = S(SN0 ) 1 N0 :
306
Cosmography
of Wiener filtering and power spectrum estimation. He gave the formulae for the
Wiener filter in the signal-to-noise eigenmode basis and showed how to compute
minimum variance power spectrum estimates from the Wiener-filtered modes.
9.3.1 Filtering in spherical basis
The Wiener filter method can be applied both to the redshift surveys and to the
peculiar velocity surveys. The usual approach in the case of redshift surveys is to
work in spherical coordinates and to use the FourierBessel expansion to describe
the density contrast. Detailed descriptions of that procedure can be found in the
first application of that method to the 1.2 Jy IRAS redshift survey by Fisher et al.
(1994) and in the Wiener reconstruction of the PSCz survey by Schmoldt et al.
(1999). The FourierBessel expansion for the density contrast is
(r; br) =
lmn
(9.12)
(r)
where b is the direction in the space (angular coordinates), Ylm b are spherical harmonics, jl x are spherical Bessel functions, and the discrete set of radial
wavenumbers kln is determined by the boundary conditions. There exist different choices for that (see Fisher et al. 1994). The most popular choice is to ignore
;r > R
possible density fluctuations after the boundary of the survey
and to demand that the logarithmic derivative of the gravitational potential should
be continuous at the boundary. In this case, the wavenumbers kln are the zeros of
jl 1 kR .
The inverse transform is
()
(r) = 0
( )
lmn = Cln
? (br);
(r; br)jl (kln r)Ylm
V
where Cln are the normalization constants, defined as:
Z R
K 0C 1
jl kln r jl kln0 r r2 dr nn
ln
) (
(9.13)
Cln1 =
R3
jl (kln R) 2 :
There exist different conventions for the FourierBessel transform. Fisher et al.
(1994) include the coefficients Cln in the sum for the direct transform (9.12) and
omit them in (9.13); we follow the convention adopted in Schmoldt et al. (1999).
The amplitudes lmn now describe the real density field. Limiting the index
range in the sum amounts to smoothing, as higher-order angular and radial modes
are usually noisy. The usual method is to choose l to limit the angular resolution
317
2002 by CRC Press LLC
Wiener filtering
307
slmn = Cln
? (br )j (k s );
w(si )Ylm
i l ln i
where the summation is over all the objects i, si and bi are their redshift distance
and the position on the sky, respectively, and w s is the weight function. The
radial velocity correction assumes the role of the response function above, as
the relation between the contrast amplitudes in real and redshift space can be
written as
X
s
lmn
(9.14)
Zlnn0 lmn0 ;
n0
where the redshift space density contrast is computed as
()
=
1
s = s
K K O :
lmn
lmn l0 m0 n
The term On is the monopole term (the transform of the constant 1):
On =
4R3
j1 (k0n R)
:
k0n R
Using the radial velocity distortion operator (8.46), the response matrix can be
written as
Z R
ll
K0
jl kln0 r
Zlnn0 nn
jl kln r
2 0 r2
kln
0
jl0 kln0 r d r 2
r dr;
(9.15)
kln0 r d r
( + 1) 1 (
+ ( ) lnln ( )
)
)+
where we have used the inverse weighting, following Fisher et al. (1994), w(r) =
1=(r), (r) is the radial selection function of the survey. Generalization of these
formulae for a general weighting function can be found in Schmoldt et al. (1999).
9.3.2 Density interpolation
We saw in the section on the estimation of power spectra that radial and angular
modes are mixed due to the limited angular window of the survey. The reconstruction ideology uses a different approach extrapolating the observed density
distribution over the full sky. Predicting the structures that we cannot yet observe
is logically also one of the goals of reconstruction.
318
308
Cosmography
The first studies manually extrapolated the density into the masked-out regions.
Later approaches use the minimum-variance method to do that. Schmoldt et al.
represent the observed redshift space density distribution by an ansatz
s (s; br) =
lmn
that separates radial and angular modes at the expense of postulating the radial dependence. The expansion coefficients are estimated by minimizing the integrated
variance
2 =
b s
(r)
(s; br)
lmn
(r)
(r ) (
(r)
(r) (
? b j k s s2 d
b Ylm
0 0n
r2 d
dr;
d
is the solid angle element.
ds = C0n1 sl0m0 n
l0 m0
? (br) d
:
(br)Yl m (br)Ylm
0
(9.16)
Inverting (9.16), we find the redshift space density amplitudes slmn and supplement the catalog by filling in the masked regions (generating galaxies by an inhomogeneous Poisson process). Fig. 9.2 shows an example (from Schmoldt et al.
1999) that compares the galactic x-z coordinate plane of the PSCz redshift survey
before and after filling the masked zones. It can be seen that the procedure interpolates the density distribution, taking into account the structure of the survey
data.
Saunders and Ballinger (2000) generalized this approach for three-dimensional
interpolation of the density field into masked regions. They assumed that the density field is log-normally distributed and the observational survey is randomly
sampled for that field. The density logarithm can be expanded in the Fourier
Bessel basis,
"
#
s (x) = exp
(x)
an fn
;
n
where n is the collective index (lmn). Neglecting correlations, they write the loglikelihood L as
XX
L
L
an fn m
m n
(the index m denotes summation over galaxies) and maximize it together with the
constraint that the total number of galaxies should be that observed. This gives
the system of equations
@L
@an
= ln =
(x )
"
X
fn (xm )
(x)fn (x)exp
319
an fn (xm ) d3 x;
Wiener filtering
309
Figure 9.2 Projections of a slice of the PCSz survey galaxies to the galactic x-z plane (the
; b ) is horizontal,
y range is
; h 1 Mpc). The x-axis (directed toward l
1
) is vertical, units are in h Mpc. The left panel shows the raw
the z -axis (toward b
data, the right panel shows the structure after filling in the masked zones, the horizontal
wedge of the Zone of Avoidance, and the high-altitude mask strip. The added galaxies are
shown by light gray. (Reproduced, with permission, from Schmoldt et al. 1999, Astronom.
J., 118, 11461160. AAS.)
( 20 20)
= 90
=0 =0
which can be solved for the full set of (redshift space) density amplitudes. They
test the procedure on simulated catalogs and find that it works well.
9.3.3 Wiener reconstruction
Equation (9.14) shows that for a full-sky survey the response matrix ( in our
case) can be solved separately for all angular modes. For that we need the expressions for the noise covariance matrix
Z R
Nlnn0 Cln Cln0
w r jl kln r jl kln0 r r2 dr
Slmnn
Cln Cln
()
) (
Z RZ R
0
() (
? (br )Y (br )
jl (kln r1 )jl (kln r2 )Ylm
1 lm 2
0
where r is the correlation function of the density field. Writing that as a Fourier
transform of the power spectrum
(r1
r2 ) =
P (k)e ik(r1
320
2 dk
(2)3
r2 ) k
310
Cosmography
eikr = 4
lm
? (br)Y (k
il jl (kr)Ylm
lm b)
we get
Slnn
Z 1
Z R
2
2
= Cln Cln
P (k)k dk
jl (kln r1 )jl (kr1 )r12 dr1
0
Z R
2k2 D (k
kln ):
Another integral over r gives the normalization constant, and finally we get
K 0C P k :
Slnn0 nn
ln
ln
( )
This shows that for very narrow windows the signal covariance matrix is diagonal with elements proportional to the values of the power spectrum at discrete
wavenumbers. None of the matrices ; ; depends on m.
In order to calculate the covariance matrix we have to choose the prior power
spectrum. This spectrum also can be obtained by a maximum likelihood fit of the
same data, but that is not strictly necessary.
Applying now the Wiener filter (9.10) with the response matrix
, we
reconstruct the density field in real space:
ZNS
R=Z
Knowing the density field, we can also reconstruct the gravitational potential
by
= 23
M H02 b
(the we observe is the galaxy overdensity, connected with the mass overdensity
M by bM , where b is the bias factor). Expanding in the FourierBessel
series and using the fact that jl kr Ylm b is an eigenfunction of the Laplacian,
( ) (r )
[jl (kr)Ylm (br)] = k2 jl (kr)Ylm (br);
we find that the FourierBessel amplitudes of the gravitational potential are simply
M H02 lmn
(9.17)
lmn
2 :
b kln
= 3
2
321
2002 by CRC Press LLC
Wiener filtering
311
M)
:
v = 32 f
(
M H0
u(r; br) = H0
jl0 (kln r)
Ylm (br)lmn :
lmn kln
X
The formulae for the transverse components are more complex; they can be found
in Fisher et al. (1994).
9.3.4 Maps
We illustrate the Wiener reconstruction procedure in Fig. 9.3. It shows the steps in
density reconstruction of the IRAS 1.2 Jy survey by Fisher et al. (1994). The upper left panel shows the raw redshift space density field, smoothed by a Gaussian
kernel with the width growing toward larger radii. The upper right panel shows
the density obtained by the FourierBessel expansion of redshift data. Here the
smoothing is smaller and the density field is noisier. The lower left panel shows
the result of the Wiener filtering in redshift space. We see that the Wiener filter
has suppressed high-frequency noise, with stronger smoothing in the outskirts of
the survey, where the shot noise is stronger. The lower right panel shows the final result, the application of the Wiener filter with correction for radial velocity
distortions. The main differences between the last two density distributions are
higher density amplitudes and more anisotropic structures in redshift space. The
overall reconstructed density distribution is similar to that obtained by the POTENT method and illustrated above, but more detailed, due to the much larger
size of the survey.
Fig. 9.4 shows the other reconstructed fields along with the density field (upper
left panel). The gravitational potential in the upper right panel is much smoother
than the density. Its main features are the large positive region, corresponding to
the Sculptor void, and two deep minima at the regions of the Great Attractor and
the PerseusPisces superclusters. The radial velocity field is shown in the lower
left panel and the full peculiar velocity field is shown in the lower right panel.
The main features of the velocity field are the infall patterns toward the two main
density maxima.
Comparison of this picture with the POTENT solution (Fig. 9.1) shows that
while there are slight differences in the density distributions, the velocity field
predicted on the basis of the density distribution does not describe the divergencefree component of the flow through the region revealed by peculiar velocity measurements. It can be caused by an insufficient depth of the 1.2 Jy IRAS survey
the velocity field feels gravitational (tidal) forces from larger distances. This
assumption is proved by the reconstruction of the velocity field on the basis of
322
2002 by CRC Press LLC
312
Cosmography
Figure 9.3 Reconstruction of the 1.2 Jy IRAS density field in the supergalactic plane. The
: , the heavy contour denotes
, and dashed
density contour interval is
lines show negative contours. Panel a) shows the raw redshift space density distribution,
smoothed with a variable width Gaussian filter. Panel b) shows the redshift space density
in the FourierBessel expansion with lmax
and kn R <
. Panel c) gives the
redshift space density after smoothing with the Wiener filter, and panel d) shows the Wiener
reconstructed real space density field. (Reproduced, with permission, from Fisher et al.
1994, Mon. Not. R. Astr. Soc., 272, 885910. Blackwell Science Ltd.)
= 05
=0
= 15
100
the PSCz redshift survey (Schmoldt et al. 1999). This survey has about two times
more galaxies than the 1.2 Jy survey and is about twice as deep. This field is
shown in Fig. 9.5.
We see that the velocity pattern in the central region is already more similar to
the POTENT solution. The backfall into the Great Attractor region (the Centaurus
and PavoIndus superclusters) is clearly seen (although the authors do not yet
consider it firmly established) and the Shapley supercluster has emerged in the
velocity map.
323
2002 by CRC Press LLC
Wiener filtering
313
Figure 9.4 Reconstruction of the 1.2 Jy IRAS density, potential, and velocity fields for the
supergalactic plane. Panel a) shows the reconstructed density field, and panel b) shows
. Panel c) shows the radial velocity field, with
the gravitational potential, assuming
open dots showing negative velocities and closed dots indicating positive velocities. Panel
d) shows the projection of the full peculiar velocity field on the plane. (Reproduced, with
permission, from Fisher et al. 1994, Mon. Not. R. Astr. Soc., 272, 885910. Blackwell
Science Ltd.)
=1
314
Cosmography
Figure 9.5 Reconstruction of the peculiar velocity field for the PSCz redshift survey. The
figure shows the projection of the velocities on the supergalactic SGX SGZ plane (the
horizontal axis is SGX , the vertical axis is SGZ ). Coordinate units are h 1 Mpc. The
locations of the main superclusters are marked. Double contours show the zero density
contrast level. (Reproduced, with permission, from Schmoldt et al. 1999, Astronom. J.,
118, 11461160. AAS.)
u = br v + n:
Here is the collection of the observed radial peculiar velocities, are the true
peculiar velocities, b is the direction of the radius-vector, and are the instrumental errors of measurement of radial velocities. In components, this relation
325
2002 by CRC Press LLC
Wiener filtering
315
Figure 9.6 Reconstruction of the real space density for the PSCz redshift survey. The figure
shows the three-dimensional density distribution in supergalactic coordinates; the coordinate units are h 1 Mpc. The locations of the known superclusters are marked. (Reproduced, with permission, from Schmoldt et al. 1999, Astronom. J., 118, 11461160. AAS.)
is
ui =
rbi vi + ni ;
where the lower (Latin) indices enumerate galaxies and the upper (Greek) indices
that may have values from 1 to 3, the vector components. To avoid confusion, we
shall write explicitly below all sums over indices.
The Wiener estimator for the velocity can now be written as
vbk =
i;j
hvk ui iDij1 uj =
Dij = hui uj i =
ij
Vki Dij 1 uj ;
(9.18)
The linear approximation relation between velocity and density contrast (9.3) allows us to write a similar Wiener estimator for the density contrast:
bk =
i;j
hk ui iDij1 uj =
326
ij
Tki Dij 1 uj :
(9.19)
316
Cosmography
The velocity covariance tensor is the main object of study in turbulence (classical turbulence describes incompressible liquids, so velocity is the only random
field there). The velocity covariance tensor can be written as (Monin and Yaglom
1975):
K
hvi vj i Sij
(9.20)
? rij
? rij rbij rbij ;
k rij
== ( ) + ( ) ( )
where k (r) and ? (r) are the radial and transverse velocity correlation functions, rij = ri rj and rij = jrij j.
k (r) =
(9.21)
? (r) =
(9.22)
where j0 (x) and j1 (x) are the spherical Bessel functions. Using (9.20), the expression for the data covariance matrix simplifies to
Vij =
(r ) =
H0 f (
M ) 1
22 0 P (k)j1 (kr)k dk:
Z
(9.23)
Additional density and velocity smoothing can be done by multiplying the power
spectrum P k in formulae (9.21 and 9.23) by a squared Fourier transform of a
smoothing filter, for Gaussian smoothing of width R, for example, by
()
exp( k2 R2):
Zaroubi, Hoffman, and Dekel (1999) applied the Wiener filter to the Mark III
peculiar velocity catalog. The comparison of the Wiener-filtered peculiar velocity
field from Mark III and the fields predicted by the IRAS 1.2 Jy survey are compared in Fig. 9.7. The density and velocity fields (upper panels, the IRAS field on
the left, the Mark III field on the right) are only roughly similar. The difference
also might be due to the complex nonlinear reconstruction procedure of the IRAS
327
Wiener filtering
317
Figure 9.7 Comparison of the density and velocity fields obtained by Wiener filtering of a
redshift catalog (IRAS 1.2 Jy survey) and a peculiar velocity catalog (Mark III catalog).
The upper panels show the density and velocity distributions in the supergalactic SGX
SGY plane, the IRAS 1.2 Jy survey on the left, the Mark III on the right. The middle
panels show the flow lines for these catalogs (the lengths of the lines are proportional
to the velocity amplitude). The lower panels show the different components of the Mark
III velocity field, the divergent component responsible for the local density structure on
the left, the remaining tidal component on the right. (Reproduced, with permission, from
Zaroubi, Hoffman, and Dekel 1999, Astrophys. J., 520, 413425. AAS.)
328
2002 by CRC Press LLC
318
Cosmography
PSCZ
50
-50
-50
50
Figure 9.8 Wiener reconstruction of the density and velocity distributions of the PSCz redshift survey in the nearby region of the supergalactic plane. (Reproduced, with permission,
from Schmoldt et al. 1999, Astronom. J., 118, 11461160. AAS.)
1.2 Jy density field. Comparison with the straight Wiener-filtered PSCz density
map (Fig. 9.8) shows a better agreement.
More interesting is the comparison of the velocity fields in the middle panels,
which look very different. The lower panels show a decomposition of the Mark
III velocity field into its divergent and tidal components. As expected, the divergent fields are similar, but the redshift survey reconstruction has not recovered
the tidal field. Even the large-scale reconstruction in Fig. 9.5 does not show that
component, so it must be due to more distant regions of space.
Another problem where the effect of the large-scale velocity field is clearly
seen is the monopole problem prediction of the Local Group velocity. This can
be estimated by different methods, and as the surveys get deeper, the direction
of that velocity changes. We know the final answer; it should be the same as our
velocity with respect to the CMB background, but thus far differences remain.
329
2002 by CRC Press LLC
Constrained fields
319
(k)
(k) =
FW (k) =
(k) (k)
hje(k)j2 i :
hje(k)j2 i + hjnej2 i
Tildes denote Fourier amplitudes: e are the true Fourier amplitudes, eO are the
e are the
raw observational estimates, eW are the Wiener filter estimates, and n
noise amplitudes.
The variance of the Wiener filter estimate is
hFW2 O2 i
2
hj
j2 i
= hjj2i + hjnj2 i hjj2i + hjnj2 i
hj
j2 i
= hjj2i + hjnj2 i hjj2i:
This is always smaller than the variance of the true field. Also, in the outer regions of surveys, where shot noise may dominate, the Wiener-filtered signal will
approach zero. This means that the reconstruction of the true field is not uniform
over the survey volume, which makes comparison with other methods difficult.
Two approaches are used to change this property of Wiener filtering. One approach, the power-preserving filter, proposed by A. Yahil, is described and applied
to galaxy surveys by Sigad et al. (1998). This filter is
FY (k) =
FW (k):
Calculating the variance of the estimate as above we find that hFY2 jeO j2 i hjej2 i,
so the variance of the field is preserved. The filter is almost optimal, with the mean
square residual only slightly larger than that for the Wiener filter. The mean field
will also approach zero in the regions dominated by noise, but more slowly.
Another approach is that of restoring the small-scale features of the true field.
In order to apply the filters, we had to choose the prior power spectrum, so we
already have used knowledge about the true field. Statistical homogeneity of the
reconstructed field is used to produce random realizations of the field that are
consistent both with the prior and the data.
Let us develop the signal-data model
d = Rs + n
(9.24)
further and suppose that the true signal s and the noise n are Gaussian homo-
320
Cosmography
1 p exp 1 sS 1sy ;
(9.25)
2
(2)m=2 jSj
and a similar formula for fn (n) with the noise covariance matrix N. The probabilfs (s) =
ity density (9.25) describes a free field, with the statistical properties determined
in our case by the prior power spectrum. But our realizations have to satisfy the
M observational constraints (9.24). The conditional probability density for such
a field is, by Bayes theorem,
f (s)f (djs)
:
f (sjd) = s
fd (d)
(d s)
n = d Rs
1 syS 1 s + (d Rs)yN 1 (d Rs) :
f (sjd) / exp
2
f (sjd) / exp
where
1 uyQ 1 u ;
2
(9.26)
u = s SRy(RSRy + N) 1d
Q = (S 1 + RyN 1R) 1:
and
(9.27)
Although these expressions might seem a little complicated, they show that the
constrained field is also a Gaussian field (but not homogeneous any more) with a
mean value
y
y
1
(9.28)
and the covariance matrix . Because (9.28) coincides with the Wiener filter
estimate, it shows that the Wiener filter gives the mean value of the conditional
signal distribution. For the Gaussian distribution this is also the most probable
value for the signal vector.
We can also see from above that the distribution of the residual field depends
only on the covariance matrices and and on the form of the response operator , and not on the values of the measurements themselves ( is a zero-mean
Gaussian field). This leads us to the procedure of constructing constrained realizations of the field , first proposed by Hoffman and Ribak (1991). We start with
known ; ; , and the response operator . The next steps are as follows:
1. Generate a realization of the Gaussian random field 0 with the prior power
0 for this particular
spectrum and find the values of the measurements 0
realization. These values will certainly differ from the measured values we
have.
u = SR (RSR + N) d
Q
S
dSN
331
2002 by CRC Press LLC
s
d = Rs
Constrained fields
321
2. Find the residual field using (9.27); this field will also be the residual field
of the final constrained realization.
3. Find the mean value of the constrained realization using (9.28) and add to it
the residual field.
This sums up to the formula
anyway, generating constrained realizations does not require extra effort. We have
described methods of generation of Gaussian random fields in Chapter 7.
An example of a Wiener-filtered mean field and its constrained realizations,
given in Fig. 9.9, describes the density and velocity distribution of the Mark
III catalog. The upper left panel shows the reconstruction (the density field is
smoothed by an extra Gaussian of width h 1 Mpc). The upper right panel shows
the signalnoise ratio of the density field, defined as
S=N
= jW ((xx))j ;
W
where W is the predicted density contrast and W is the rms residual from (9.11).
The first contour of S=N starts at unity, and we see that clear structure is seen only
in the two large supercluster regions and in the Sculptor void. The middle and
lower panels show the constrained realizations of the same density and velocity
fields. We see that as the main features remain approximately same, the Wiener
filter estimate still leaves considerable freedom for small-scale details.
Nevertheless, the small-scale structure is affected by the constraints imposed by
the large-scale environment. An example of an analysis of such an influence is the
work of van de Weygaert and Hoffman (2000) on the peculiarities of the cosmic
flow in our neighborhood. It is a long-known and uncomfortable fact that the flow
velocity here is appreciably less than the predictions of N -body models. van de
Weygaert and Hoffman show that it is possible to build constrained realizations
of a local cold flow, conditioned by the strong tidal shear in our neighborhood.
9.4.1 Constrained realizations for models
The derivation above used the fact that the distribution of measurements conditioned by the field coincided with the noise distribution. However, the formalism
of constrained realizations of Gaussian fields also can be applied to the case where
there is no noise, and the operator describes constraints we wish to apply to a
field (e.g., for simulation). In fact, generation of constrained initial fields for numerical simulations was the motivation to develop the formalism (Bertschinger
1987). In that case we can write
1
1
f (djs) / exp
2 dC d ;
332
2002 by CRC Press LLC
322
Cosmography
Figure 9.9 Wiener reconstruction of the Mark III catalog. The upper left panel shows the
reconstructed density and velocity fields; the upper right panel shows their signalnoise
ratio (with the contour spacing one). The density has been smoothed by a Gaussian of
width h 1 Mpc. The middle and lower panels show different constrained realizations of
the data in the upper left panel: the middle panels show only the density, the lower panels
show both the density and velocity fields. The spacing of density contours is 0.15 in the
upper and middle panels and 0.3 in the lower panels. (Reproduced, with permission, from
Zaroubi, Hoffman, and Dekel 1999, Astrophys. J., 520, 413425. AAS.)
333
2002 by CRC Press LLC
Constrained fields
323
because the constraints are linear, and the covariance matrix of constraints
given the field is
C = RySR:
Using that, the formula for f (sjd) will coincide with the formula (9.26) if we
omit there the terms containing N:
s = s0 + SRyC 1(d d0):
For a true continuous representation of a random field linear constraints are
linear functionals of the field, including values of the field and its derivatives at
selected points. A detailed description of constructing such realizations and examples of their application for generating the initial fields for numerical simulations
are given by van de Weygaert and Bertschinger (1996). Linear constraints can be
with a kernel Hi
imposed as convolutions of the field s
i :
(x)
(x; x )
di = s(x)Hi (x; xi ) d3 x:
Z
In this case the constrained field can most easily be found from its Fourier transform se :
(k)
(k)
d0j );
()
where se0
are the Fourier amplitudes of the unconstrained realization, P k is
ei
is the Fourier transform of the convolution
the power spectrum of the field, H
kernel, dj is the value of the j -th constraint, and d0j is the value of the same constraint for the unconstrained realization. The covariance matrix of the constraints
Cij is given by the Fourier integral
(k)
Cij =
d3 k
(2)3 :
The main features of constrained realizations are illustrated by a simple example in Fig. 9.10. This figure shows the density profiles for slices of constant
height z through a constrained density realization. The smooth density distribution (constraints) is represented by three ellipsoids in the upper left panel. While
the constraints for the observational catalogs usually take the values of the field at
selected points, an ellipsoidal density peak has to be determined by 10 constraints
(its height, position, orientation, and shape; van de Weygaert and Bertschinger
1996). Specifying the velocity field will add another 10 constraints per peak. We
have not done that; the constrained realizations in Fig. 9.10 satisfy 30 constraints.
The three remaining panels show different z -slices (denoted by the slice number) of the same constrained realization. The field was generated in a 3 cube, so
these slices are all close to the center of the cube. The high density constraint in
the central slice has effectively suppressed fluctuations at the locations of the two
main peaks, but the smaller peak is severely distorted. Traces of regular structure
disappear quickly with increasing distance from the center. In the three centers
32
334
2002 by CRC Press LLC
324
Cosmography
16
16
15
14
Figure 9.10 Constrained realizations of a Gaussian density field (the height of the surface
gives the density contrast). The constraints are specified as the three ellipsoids in the upper left panel. The remaining panels show the realizations in different z -slices of a 3
computational cube; the slice numbers are shown.
32
the dispersion of the field is zero, but it grows fast outward and soon achieves
its value for the free field. The outer regions of the cube in the upper left panel
show a picture typical for a Wiener-filtered field, and the realizations give there a
uniform random density field.
The realizations in Fig. 9.10 were computed using Bertschingers COSMICS
package of generation of initial conditions for cosmological simulations. It can be
found at the (COSMICS) Web page.
A useful generalization of the above formalism was developed by Sheth (1995).
He showed that we can use similar methods to construct constrained realizations
of random fields that can be obtained by transformations of
p a Gaussian field. This
class includes log-normal fields, 2n fields, and Rayleigh ( 2n ) fields. Such fields
are frequently used as statistical models in cosmology.
9.5 Time machines
Reconstruction of physical fields in our neighborhood can be developed further,
asking how they looked in the past. That knowledge would allow us to check
the basic assumptions of the structure formation theories and enable us simply
335
2002 by CRC Press LLC
Time machines
325
to satisfy our curiosity about the past. We can never observe that, because our
observations take us along the light cone; the past we see is the past of far-away
regions in space.
There are also technical requirements for that. Creating Gaussian constrained
realizations for present-day reconstructions is wrong, in principle, as the density
distribution at small scales is no longer Gaussian. The right way to add small-scale
structure is to take the present structure into the past where it should be Gaussian,
to build constrained realizations there, and to evolve these dynamically back to the
present time. The resulting fields are used to create reference catalogs to compare
with the observed data and to estimate the errors of the analysis. This approach
has been realized by Kolatt et al. (1996) and Bistolas and Hoffman (1998).
The main problem here is the choice of a good time machine that would
calculate for us the density and velocity distribution at an early time, given the
present-day data. Straight numerical N -body modeling will not work. Due to the
noise in the present-day reconstruction, there is always a mixture of the decaying
mode of perturbations, and N -body modeling would amplify it into the past, soon
losing the growing mode we are looking for. Thus, the time machines are usually based on dynamical approximations that exclude the decaying mode. There
are several such methods; a review and a comparison of these methods are given
by Narayanan and Croft (1999) (the only method published later is that of Monaco
and Efstathiou 1999).
Narayanan and Croft divide the methods into three classes. The first class starts
with the linear solution, where the density is proportional to the growth rate of
the growing mode D1 t that is a known function of time, and its higher-order
generalizations. A second approach assumes that the rank order of the density
field does not change during evolution; the highest density peaks remain the highest and the lowest densities remain the lowest. The present density distribution is
re-mapped to a Gaussian distribution and this is taken into the past with one of
the time machines of the first class. This gives a better initial density distribution
than the first methods, but cannot be used when we want to study the one-point
probability distribution of density, because it is transformed into a Gaussian by
force. The third class of methods is based on the least action principle, and, in our
view, is the most interesting; we shall describe that approach below.
Peebles (1989) proposed using the least action principle to trace galaxy orbits
back in time. He applied that method to nearby galaxies; Croft and Gaztanaga
(1997) reformulated the method for cosmological reconstruction of large galaxy
catalogs.
In the Hamiltonian formulation of dynamics the properties of a mechanical
system are described by a functional called action:
()
S=
Z t
L(x; x_ ) dt;
336
2002 by CRC Press LLC
326
Cosmography
where the Lagrangian L is the difference of the kinetic energy K and the potential
energy W ,
L = K W:
The equations of motion of a system of particles can be found from the requirement that the action has to be stationary (minimal) when the particle trajectories
are varied. Thus, the requirement of least action is equivalent to finding all true
paths of particles.
For a system of points in an expanding background the energies are (Peebles
1989):
K = 12
mi a2 jx_ i j2 ;
(9.29)
x_
mi jxi j2 ;
(9.30)
where mi ; i , and i are the masses, coordinates, and velocities of the particles, a
is the dimensionless scale factor, G is the gravitational constant, and is the mean
density of the universe. The dependence of these energies on the scale factor is
due to the expansion and the second term in the potential energy is the relativistic
curvature term, the energy of the mean smooth distribution of matter. The kinetic
and potential energies are connected by the cosmic energy conservation equation
(Peebles 1980, x 24)
d
[a(K + W )] =
dt
Ka_ :
We can write the Zeldovich approximation for particle paths and velocities as:
(9.31)
(9.32)
x_ (0)
where we have supposed that the particle velocities in the past were small, i
j j2 =
mi j i j2 :
2Z a
j
j
W = K 2a
a2 D_ 2 da
o
337
2002 by CRC Press LLC
Time machines
327
S = j j2
Z 1
D_ 2 a2 +
1 Z a D_ 2 a2 da da :
2a 0
a_
10
is = Pi jj
(i is the real space displacement), where the redshift space projection tensor is
i = Pij 1 js ;
338
(9.33)
328
Cosmography
Figure 9.11 Gen-PIZA-reconstructed PSCz velocity field in the supergalactic plane (a slice
20h 1 Mpc thick). The arrows show the galaxy displacements, starting at the initial position and ending in the PSCz real-space position. (Reproduced, with permission, from
Valentine, Saunders, and Taylor 2000, Mon. Not. R. Astr. Soc., 319, L13L17. Blackwell
Science Ltd.)
( is the radial vector in redshift space). The formula (9.33) gives an expression
for the square of the real-space displacement
i2
=( )
is 2
1
1 (1 + )2 jsi brj2 :
There are a few additional finesses. First, in the case of a magnitude-limited survey the masses assigned to galaxies (and PIZA particles) are inversely proportional to the selection function and low-mass pairs could get large displacements,
so additional constraints must be used to eliminate large displacements for smallmass pairs. Second, as Taylor and Valentine (1999) have shown, the velocity
reconstruction is less noisy when the Local Group velocity (the dipole term) is
subtracted. The resulting reconstruction of the PSCz velocity field is shown in
Fig. 9.11.
339
2002 by CRC Press LLC
Gravitational lensing
329
If we compare this velocity field with the Wiener-filtered direct velocity measurements of the Mark III catalog (Fig. 9.7) we see that the two velocity fields are
rather close. The PIZA reconstruction has captured well the divergent component
of the flow and, to some extent, the tidal component.
9.6 Gravitational lensing
Gravitational lensing is a powerful tool that allows us to measure directly the total
matter content in the universe. This is a very broad and well-developed topic, with
applications that are closer to image reconstruction than to three-dimensional spatial statistics. Although the theory of application of gravitational lensing for studying the large-scale structure of the universe was developed in the early 1990s, the
first observational detections appeared only recently. Here we present only a brief
introduction to the subject.
9.6.1 Physics of gravitational lensing
General reviews of gravitational lensing can be found in Schneider, Ehlers, and
Falco (1992) and Narayan and Bartelmann (1999). Our attention will be focused
on weak lensing and we follow below mainly the review by Bartelmann and
Schneider (2001).
Gravitational lensing is based on the well-known fact that light is deflected by
b for a
gravitating masses, predicted by general relativity. The deflection angle
light ray passing at a distance (impact parameter) from a point mass (star) is
4GM ;
b =
c2
where M is the mass of the star and c is the speed of light. If the mass distribution
is extended, the total deflection is given by
0 dm = 4G Z d2 0 Z 0 (0 ; z) dz;
j 0 j2
c2
j 0 j2
b is the two-dimensional deflection angle on the sky, is the corresponding
where
b = 4cG2
impact parameter vector, and z is distance from the observer. This expression
is obtained using the approximation of small deflection angles, which is almost
always satisfied. Introducing the surface (projected) mass density
() =
(; z ) dz;
b = 4cG2
0 (0 ) d2 0 :
j 0 j2
340
330
Cosmography
Figure 9.12 Geometry of an extended gravitational lens system. (Reproduced, with permission, from Bartelmann and Schneider 2001, Phys. Rep., 340, 291472. Elsevier Science.)
Let us now write the lens equation that relates the true position of the source
with its apparent position on the sky. Fig. 9.12 shows the geometry of a gravitational lens. The position of the source in the source plane can be written as
Ds
=D
Dds b ();
d
where the distances D have been defined in the figure. Replacing the position
vectors and by the corresponding angles and , we get
introducing the scaled deflection angle . As the lenses and sources are usually
far away, and the distances above are used to describe angles, these distances are
cosmological angular diameter distances (Chapter 2).
The dimensionless surface mass density (or convergence) is defined as
() =
(Dd ) ;
cr
341
2002 by CRC Press LLC
Gravitational lensing
331
where
Ds
c
cr = 4G
Dd Dds
is the critical surface mass density, which divides the occurrences of strong
(multiple images) and weak lensing. The scaled deflection angle is now
Z
0
() = 1 (0 ) j 0j2 d2 0 :
The last equation shows that the deflection angle is a gradient of the deflection potential , r , where the deflection potential is given by the two-dimensional
Poisson equation
() = 2():
Gravitational deflection will change apparent positions of light sources and will
distort the images of extended sources (galaxies). Locally the sourceimage mapping can be described by
d
d;
(9.34)
where the Jacobian is
=A
2 ( )
= ij @@ @
= 1
2
1
A() =
i j
and the shear components
1 and
2 are given by
@i
@j
1 =
1
2(
;11
;22 );
2
k +
1
2 = ;12 :
(9.35)
The shear is a trace-free part of the symmetric matrix and can be represented by
a complex number
= 1 + i 2 :
exp(2 )
() =
where
2 2 2i
D() = 2 1 4 1 2
jj
meaning of the convergence and
(9.36)
To clarify the
the shear
, we write the
linearized mapping (9.34) in the neighborhood of the image point , which corresponds to the source point 0 , as
I () = Is 0 + A( 0 ) :
342
2002 by CRC Press LLC
332
Cosmography
Figure 9.13 The central part of the Abell 2218 galaxy cluster, with strong gravitational
lensing effects. (Reproduced, with permission, from Kneib et al. 1996, Astrophys. J., 471,
643656. AAS.)
Here I represents the surface brightnesses of the image and the source, respectively. This mapping shows that a circular source of a unit radius will be transformed to an ellipse with semiaxes
a=
j j 1 ; b = 1 + j j 1 :
The total fluxes from the image and the (unlensed) source are proportional to
the areas of the image and the source, and their ratio is the magnification 0 :
=
1
det A = (1
)2
( )
j j2
The regions in the image plane, where the transformation is singular and the mag, are called critical curves. Near critical
nification is formally infinite,
curves the magnification and distortion of images are great.
Fig. 9.13 shows the Hubble Space Telescope image of the well-known lensing
cluster Abell 2218, where many strongly deformed images of background galaxies (arclets) are seen, together with a number of less deformed images. Such a
map can be used to restore the projected mass distribution of the far-away clush 1 Mpc). Because lensing is strong, several critical curves
ter (its distance is
and multiple images can be identified here (Kneib et al. 1996).
det A = 0
525
Gravitational lensing
333
The shape of a galaxy image can be described by the tensor of brightness moments Mij :
Mij =
W ()I ()i j d2 ;
where the angles are calculated in a coordinate system centered on the image
and W is a normalized weight function. The complex ellipticity e is defined as
()
e=
M11 + M22
(M11
M22 + 2iM12 )
The lens mapping (9.34) induces a relation between the source moments
the image moments :
T;
s
Ms and
M = AMA
which gives the relation between the complex ellipticities of the source es and
image e, respectively:
e g g2 e?
es
(9.37)
jgj2 < ge? :
= 1 + 2 +2 ( )
g() =
()
:
()
Taking the expectation value of (9.37), the reduced shear can be expressed via the
measured average image ellipticities. For a weak lensing regime, when ,
j
j , and jgj , we get from (9.37) the relation
e es g es
:
+2 = +2
hj ji = 21 hei:
=
for which
1 + (1 jej2 )1=2
hj ji = hi:
Thus, we can estimate the shear by locally averaging the complex ellipticities of
galaxy images. Once we know the shear distribution in the image plane
,
we can, in principle, reconstruct the projected surface density by inverting the
convergenceshear relation (Kaiser and Squires 1993):
()
Z
?
1
<
D ( 0 )
(0 ) d2 0 + 0 :
() =
344
334
Cosmography
Another approach is to use maximum-likelihood and Bayesian methods for reconstruction. A list of references and application of a maximum entropy method
can be found in Bridle et al. (1998).
9.6.3 Cosmic shear
Reconstruction of selected lensing mass distributions has been feasible for several years. Another application of gravitational lensing is measurement of cosmic
shear, produced by the overall effect of the large-scale gravitational fields. This
signal was first detected only in 2000.
Using a similar reasoning that leads us from the lens equation to the expression
for the convergence (see, e.g., Schneider et al. 1998), the cosmic convergence can
be written as
Z !
H02 R02
Sk ! !0 Sk !0 Sk !0 ; !0 0
d! : (9.38)
; !
M
c2
Sk !
a !0
0
) = 23
) ( ) [ ( ) ]
()
( )
()
( ) = 23
()
()
( ) ( ) [ a((!)) ]
( )=
( ) ( ( )!) d!0
is the source-averaged distance ratio Dds =Ds (see Fig. 9.12). Assuming, as usual,
that the density contrast is a Gaussian random field, the convergence field is also
a Gaussian random field, and it can be described by its power spectrum. Calculating the dependence between the convergence power spectrum and the threedimensional density contrast power spectrum is similar to calculating the relation between the projected and spatial density power spectra (see Chapter 8). The
only difference is in the weighting functions. A detailed derivation can be found
in Schneider et al. (1998) and in Bartelmann and Schneider (2001) (please note
that both these papers use the comoving coordinate w R0 ! , while we use the
dimensionless comoving coordinate ! ). The result is
Z
H04 R04 2 3 !H g2 !
K
R
P
! d!;
P K
(9.40)
c4 M 0 0 a2 ! R0 Sk !
( ) = 94
()
()
345
2002 by CRC Press LLC
( );
Gravitational lensing
335
( )
where P K is the convergence power spectrum, K is the modulus of the twodimensional wave vector, and P is the density contrast power spectrum. Its dependence on ! stresses the fact that this spectrum evolves with time along the
light cone.
Using the relation between the convergence and shear (9.36), we can find the
variance of the modulus of the shear in top-hat windows of radius c :
hj j2 i = 2 2
P (K )J12 (Kc )
c 0
dK
:
K
(9.41)
This is the simplest measure of cosmic shear. Another measure, the aperture mass,
is defined as
Z
Map
U d2 ;
(9.42)
<c
where U is a zero-mean (compensated) filter. Schneider et al. (1998) showed
that the aperture mass variance is given by the convergence power spectrum:
() ()
()
288 P (K )J 2 (Kc) dK :
hMap2 i =
4
4
K
c 0
Z
()
Here and above we have used the standard notation for the Bessel functions Jn x .
It is useful to define for each galaxy the tangential and radial shears
t and
r with
respect to the center of the aperture as
=
=
1 cos(2)
2 sin(2);
(9.43)
2 cos(2) +
1 sin(2);
(9.44)
where is the angle between the x-axis and the line from the aperture center to
t
r
the galaxy. Now we can express the aperture mass by the tangential shear:
Map =
()
<c
is given by
Z
t () Q() d2 ;
(9.45)
This can be used to check the origin of the signal; if it is due to intrinsic ellipticities
of galaxies, the integral is not zero.
As an example of a recent measurement of cosmic shear, we refer the reader
to the article by van Waerbeke et al. (2001). They measured ellipticities for about
2 i statistics. Fig. 9.14
1.2 million faint galaxies, and estimated the hj
j2 i and hMap
shows their results for the aperture mass statistics.
The upper panel shows the statistic itself, compared with predictions of different cosmological models, and the lower panel shows the radial shear check.
The definition of the radial and tangential shear (9.43) shows that the tangential
346
336
Cosmography
Figure 9.14 Aperture mass statistics for different aperture sizes . The upper panel shows
the Map statistics and different model predictions. The lower panel shows the R-mode,
which has to be zero for pure gravitational lensing. (Reproduced, with permission, from
van Waerbeke et al. 2001, Astr. Astrophys., 296, 374, 757769).
45
shear can be made radial by rotating a galaxy by . van Waerbeke et al. (2001)
call the mass aperture statistic for rotated galaxies the R-mode; in regions
where it differs from zero, the statistic is probably influenced by intrinsic galaxy
ellipticities or by instrumental effects. As we see, this region is rather limited, and
the mass aperture estimate in the upper panel truly represents the cosmic shear.
As this shear can be predicted, given the parameters of the cosmological model
and the model density contrast power spectrum, cosmic shear measurements can
be used as an independent check of these models.
Fig. 9.14 also shows that the amplitude of the signal is extremely small. Thus,
in order to estimate this value with necessary confidence, a large number of faint
galaxy images have to be processed. This requires considerable effort and also
involves careful elimination of various systematic instrumental effects.
45
347
2002 by CRC Press LLC
Gravitational lensing
337
( )=
q(; S ) =
N0
N
( )
(S ) :
= NN
(S )
1
=1
348
2002 by CRC Press LLC
=2
( )
= 1+2
CHAPTER 10
Structure statistics
10.1 Introduction
The picture that emerged from the description of the galaxy distribution in Chapter 1 is one of a network of galaxies concentrated in clusters, filaments, and walls
surrounding large empty voids. While the statistical measures described in Chapters 3 and 8 do provide a wealth of information about the scale, amplitude, and
even the nature of the deviations from a uniform distribution, they at best offer
only suggestive statistical measures for these structural patterns.
In this chapter we present the tools that have been developed to highlight one
or more of these structural features in the galaxy distribution. We start with a
recapitulation of the observational situation with respect to the structural features,
trying to define carefully which aspects should be described.
The first aspect of the density field that we want to describe is the topology of
the galaxy distribution. A particularly good measure to judge whether we have
a cellular or a sponge-like distribution of galaxies is the Euler characteristic of
a density field, characterized by the genus. The theory behind this quantity is
presented, together with a discussion of what is expected in a Gaussian density
field. The results of measurements in observational catalogs are presented. We
discuss the techniques for obtaining the smooth density field from the discrete
galaxy distribution and their influence on determination of the genus.
While the topological parameter provides some information on the nature of
the galaxy distribution, a real statistical description of the walls, filaments, and
clusters starts to emerge only with the development of structure functions. We
treat here the shape statistics introduced by Luo and Vishniac (1995), in addition
to the mathematically clean description defined by the Minkowski functionals.
Two older statistical techniques that focused in particular on the detection of
filaments in the galaxy distribution are percolation analysis and minimal spanning
trees. In particular, the latter may have the potential to be useful not only as a
descriptor in its own right, but also as a tool to estimate other statistical quantities
(Barrow 1992).
We will discuss the techniques that have been developed to identify voids in
the galaxy distribution. Wavelets have been introduced in cosmology as structure
detectors. We will review what they are and which aspects are useful for the analysis of galaxy clustering. We present different algorithms, many of them based on
the previous techniques, to find clusters within clustered point processes. Finally,
349
2002 by CRC Press LLC
340
Structure statistics
the appropriate statistical tools for studying any possible periodicity of structures
in the galaxy distribution are discussed.
10.2 Topological description
10.2.1 The theory of topological analysis: the genus
Topological analysis of the large-scale structure is how cosmologists refer to the
art of measuring the degree of connectivity of the matter distribution in the universe once the redshift surveys have been smoothed with an appropriate filter
function, as explained in Chapters 3 and 9. This is done by means of the topological genus (Gott, Dickinson, and Melott 1986; Mellot 1990; Coles 1992). The
genus of a surface G is basically
G=
(number of holes)
For example, a sphere has topological genus equal to 0, and a torus has genus +1,
while N disjoint spheres have G = (N 1).
The GaussBonnet theorem provides the relationship between the curvature of
the surface and its topological genus. According to this theorem, the integral of
the Gaussian curvature of a compact two-dimensional surface is
g( ) = N (1 2 )e =2 ;
(10.1)
where is the number of standard deviations from the mean density of the density
threshold and N is the normalization constant that depends on the power spectrum
of the smoothed density field
1
N= 2
4
where
hk2 i 3=2
3
2
3
hk2 i = RkPP(k(k)d)d3 kk :
350
2002 by CRC Press LLC
341
genus
Topological description
Figure 10.1 Several isodensity contours corresponding to a Gaussian field are shown in
the top panels. At the bottom, the genus curve calculated empirically (triangles and error
bars) is displayed
p together with the expected analytical expression (10.1). The smoothing
radius = 2! is shown in units of the side length. (Reproduced, with permission, from
Weinberg, Gott, and Melott 1987, Astrophys. J., 321, 227. AAS.)
1
3+n
N=
3
(2)2 23=2 !3
3=2
n 3;
where ! is the smoothing length of a Gaussian filter. The genus curve, corresponding to the relation (10.1), is symmetric about 0 in , or in other words it
is symmetric around the mean density, with the overdense regions and the underdense regions statistically indistinguishable. This is the Gaussian or random-phase
topology corresponding to a sponge-like surface. For j j < 1, g ( ) > 0 corresponding to a surface with many holes, multiply connected and negatively curved.
For j j > 1, g ( ) < 0; therefore, there are isolated clusters and isolated voids.
Isodensity surfaces can be labeled by means of the fraction fvol of the volume
contained in regions with density exceeding a given threshold, or by the number,
351
2002 by CRC Press LLC
342
Structure statistics
Figure 10.2 The isodensity contours for two models. On the left, eight clusters have been
superimposed over a uniform background. On the right, eight bubbles have been simulated.
As we can see the roles of the high-density regions and low-density regions have been interchanged in both models. (Reproduced, with permission, from Weinberg, Gott, and Melott
1987, Astrophys. J., 321, 227. AAS.)
vol , defined by
1
fvol = p
Z1
2
e t =2 dt:
2 vol
Note that, for a Gaussian density field, vol is equivalent to the number of standard
deviations that a given density threshold is above or below the average density,
but, in general, this is not the case. The topological analysis consists of studying
how the genus of an isodensity surface varies with fvol or vol .
Fig. 10.1 shows the isodensity contours for a Gaussian field in a box, where the
notation x% high (x% low) means that the depicted contours are encompassing
352
2002 by CRC Press LLC
Topological description
343
Figure 10.3 The genus curve as a function of fvol (top axis) and vol (bottom axis) calculated empirically for the models of Fig. 10.2. (Reproduced, with permission, from Weinberg,
Gott, and Melott 1987, Astrophys. J., 321, 227. AAS.)
the most dense (less dense) x% of the total volume of the box. The corresponding
genus curve is shown in the same figure calculated over four realizations of this
model; the triangles show the average and rms fluctuations are represented by
error bars.
If the curves of the genus are biased to the left the topology corresponds to isolated clusters superimposed on a smooth background. It is referred to as meatball topology. Finally, if the curve is biased to the right, the topology is the
Swiss-cheese kind, with empty voids surrounded by one connected high density region. Weinberg, Gott, and Melott (1987) and Melott (1990) analyzed several
point distributions by means of this formalism. In a model dominated by clusters,
6,400 points were distributed in a 323 lattice, 75% randomly and 25% on eight
clusters. A similar model but with eight bubbles was also generated. Fig. 10.2
shows some isocontours corresponding to these models and Fig. 10.3 shows the
corresponding genus curves illustrating the behavior explained above.
10.2.2 Estimation of the topology, technicalities
The first step in applying this kind of analysis to real three-dimensional galaxy
distribution is to smooth the point process to obtain a representation of the continuous density field. Some methods were presented in Section 3.3.1. A detailed
study is presented in Melott and Dominik (1993). Once the discrete point distribution has been smoothed, the genus statistic is applied to the isodensity contours.
The shape of the genus curve is clearly affected by the election of the bandwidth . Larger values of tend to oversmooth the data, creating positive genus,
while smaller values of produce undersmoothing and, thus, negative genus. This
variability is also useful when applying the genus statistic to real galaxy catalogs:
353
2002 by CRC Press LLC
344
Structure statistics
Figure 10.4 Isodensity contours of the QDOT survey. The top panel shows the contours
enclosing the regions corresponding to one third high-density volume and the bottom panel
shows the contours enclosing the regions corresponding to one tenth high-density volume.
(Reproduced, with permission, from Moore et al. 1992, Mon. Not. R. Astr. Soc., 256, 477
499. Blackwell Science Ltd.)
354
2002 by CRC Press LLC
Topological description
345
10
PSCz
= 12 h-1Mpc
-5
-10
-3
-2
-1
Figure 10.5 The genus curve of the PSCz redshift survey for a smoothing radius =
1
Mpc (solid line). The dashed line is the best Gaussian fit to the empirical curve.
(Reproduced, with permission, from Canavezes et al. 1998, Mon. Not. R. Astr. Soc., 297,
777793. Blackwell Science Ltd.)
12 h
It has become standard to show the genus curve applied to the same data set for
different values of the smoothing radius.
When a given isodensity surface is approximated by a network of polygonal
faces with the grid size much smaller than the smoothing radius, the genus of the
surface G verifies
X
(10.2)
Di = 4(1 G);
i
where
X
Di = 2
Vi
i
is the angle deficit around vertex i and Vi are the angles around the vertex. As
an example, we can approximate a sphere by a cube. It has 8 vertices. The angle
deficit at each vertex is =2 = 2 3=2, and therefore the expression (10.2)
provides a value for the genus G = 0, as expected for the sphere.
10.2.3 Topological measurements: observations
Several galaxy redshift catalogs have been analyzed thus far using the genus
curve, the QDOT catalog (Moore et al. 1992), the CfA2 catalog (Vogeley et
al. 1994), and the PSCz catalog (Canavezes et al. 1998). Fig. 10.4 shows the
isodensity contours of the QDOT survey encompassing roughly one third (left
panel) and one tenth (right panel) of the total volume. Different large scale structures, clusters, and superclusters, mentioned in Chapter 1, are highlighted in these
355
2002 by CRC Press LLC
346
Structure statistics
Mi =
and
Mij =
1X i
(x
N k k
1X i
(x
N k k
xi0 )
xi0 )(xjk
xj0 )
where the sums are taken over all galaxies in the window.
From these quantities Luo and Vishniac (1995) have defined the following
cubic structure statistics as (repeated indices in a term means summation from
1 to 3)
1
f M ii (M ij M i M j )2
(M ii )3
1
+ M ii (M kk (M j )2 )
2
+3M ij (M jk M j M k )(M ik M i M k )
3 ij ii
M (M
M k M k )(M ij M i M j )g
2
1
f+4M ii (M ij M i M j )2
=
(M ii )3
4M ii (M kk (M j )2 )
12M ij (M jk M j M k )(M ik M i M k )
356
2002 by CRC Press LLC
(10.3)
Structure functions
347
+12M ij (M ii
M k M k )(M ij
M i M j )g
(10.4)
Luo, Vishniac, and Martel (1996) have applied these statistics to galaxy redshift
surveys. They denote by hSr i the mean values of the shape statistics calculated
for ten realizations of a binomial process with the number of points equal to the
number of galaxies in the galaxy survey and within the same volume. Fig. 10.6
shows the average for all the windows of a given radius R0 , hS hSr ii, as a
function of the diameter of the window, 2R0 , for the PerseusPisces sample. It
can be appreciated that the plane shape statistic has a peak around 15 h 1 Mpc.
This is probably a consequence of the alignment of fingers-of-God. The line shape
statistic declines monotonically from 5 h 1 Mpc, indicating that the filamentary
structures are more evident at smaller scales.
357
2002 by CRC Press LLC
348
Structure statistics
H
3
A
M0 = V; M1 = ; M2 = 2 ; M3 = ;
(10.5)
8
2
4
where the quantities A, H , and of a compact convex body in K IR3 with
smooth boundary @K and principal curvature radii R1 and R2 are defined in
A=
H=
Z
@K
dS;
1
1
1
+
dS;
2 @K R1 R2
and
1
1
dS:
4 @K R1 R2
Note that in the last expression (K ) = (@K )=2 (Mecke, Buchert, and
=
V =
where
! =
!3
M
!3
=2
(1 + =2)
is the volume of the -dimensional unit sphere. The characterization of point fields
with Minkowski functionals is performed by associating to the point distribution
358
2002 by CRC Press LLC
Structure functions
349
Figure 10.7 An illustration of the Boolean grain model. The union set Ar is displayed for
different values of the diagnostic parameter r for an underlying Poisson distribution. (Reproduced, with permission, from Schmalzing, Kerscher, and Buchert 1996, in Dark Matter
in the Universe, IOS Press, Societ`a Italiana di Fisica.)
a union of convex sets in the following way. Each point of the process is decorated
with a ball of radius r; the radius used acts as a parameter (Mecke, Buchert, and
Wagner 1994). The functionals are then applied to the sets Ar = [N
i=1 Br (xi ).
If the underlying point process is a finite realization of a Poisson process, this
model is a special case of a Boolean grain model. Fig. 10.7 shows how this model
provides different sets Ar as the parameter r varies for an underlying random
point process.
We can generalize the Boolean model by replacing the underlying Poisson point
process by any general point process. This is known as the germgrain model or
coverage process (Hanisch 1981; Hall 1988; Stoyan, Kendall, and Mecke 1995).
Depending on the clustering properties of the underlying process, the sets Ar
will present different shapes, and thus different Minkowski functionals. In this
manner, Minkowski functionals represent a rather complete tool to study the clustering properties of galaxy and galaxy cluster distributions (Kerscher et al. 1997;
Kerscher et al. 1998).
359
2002 by CRC Press LLC
350
Structure statistics
Figure 10.8 The volume densities of the four Minkowski functionals, V , = 0; 1; 2; 3 for
a volume-limited sample of the CfA1 galaxy catalog with 80h 1 Mpc depth. The shaded
areas represent the rms error range of empirical calculations on 15 realizations of a Poisson distribution. (Reproduced, with permission, from Kerscher and Martnez 1998, Bull.
Int. Statist. Inst., 57-2, 363366.)
For a Poisson process, the mean volume densities of the Minkowski functionals are analytically known (Mecke and Wagner 1991; Schmalzing, Kerscher, and
Buchert 1996), and thus, they are usually used as a standard of reference. Fig. 10.8
shows the Minkowski functionals for a volume-limited sample of the CfA1 catalog, together with the Poisson values.
The edge correction for the estimation of the Minkowski functionals relies upon
the principal kinematic formula (Santalo 1976) that provides the Minkowski functional of the intersection of convex bodies (Mecke and Wagner 1991).
Minkowski functionals also have been used to characterize the isodensity surfaces of random fields in the analysis of the cosmic background radiation
(Schmalzing and Gorski 1998). This approach unifies and extends several statistical quantities presented throughout this book (Kerscher 2000). For example,
the volume density of the first Minkowski functional, V0 , is just the spherical contact distribution function, defined in Section 3.7 and the fourth Minkowski functional, V3 is the Euler characteristic, related to the genus of isodensity surfaces of
smoothed cosmological fields, as explained in Section 10.2.
360
2002 by CRC Press LLC
351
Sahni, Sathyaprakash, and Shandarin (1998) and Kerscher (2000) show how
particular ratios of the Minkowski functionals can be used to define shape finder
quantities (planarity and filamentarity), similar to the quantities introduced in
Section 10.3.1, and therefore provide a diagnostic about the presence of filaments,
sheet-like structures, and ribbons in the galaxy distribution.
10.4 Cluster and percolation analysis
The Minkowski functionals described in the previous section are meant to describe properties of individual objects, formed by clustering from the original
point set. These clusters are not always formal; they also could be physical objects. Cluster analysis was probably first used in cosmology by Turner and Gott
(1976) to objectively define groups of galaxies. In later years it has been used to
compile catalogs of superclusters (e.g., Einasto et al. 1997). Cluster analysis in
cosmology is usually called the friends-of-friends (FOF) method, after Press
and Davis (1982), although other techniques are also available (see Section 10.7).
The main problem in cosmological cluster analysis is the choice of a right
neighborhood radius (the radius of the ball built on a sample point) R. This gives
us the classification of points (galaxies) into clusters and allows us to find their
geometrical properties by Minkowski functionals. This is not an easy choice in
practice, as the observed clustering is hierarchical with a continuous range of
scales.
We can turn this deficiency of the method into an advantage, considering the
variation of the cluster properties with the neighborhood radius. Let us study the
process of growth of clusters, substituting the neighborhood radius for time. For
small radii the unions of the balls (ball clusters) are more or less isolated and their
total volumes are small. As the process continues, clusters grow and join together,
until at a certain radius the largest system will connect opposite boundaries of
the sample volume. If our sample was a typical region of the universe, we could
have a connected structure (cluster) flashing through the whole universe. This
moment is called percolation and the corresponding neighborhood radius is called
the percolation radius.
Percolation properties of a point sample describe the topology of the spatial
distribution the more filamentary it is, the easier it is to achieve percolation.
The usual percolation characteristic for a point sample is the average number of
points Bc in a sphere with the percolation radius Rc . Percolation does not depend
on the mean density of the sample. This can be found by numerical simulations;
for a Poisson distribution it is Bc 2:7.
Use of the percolation technique in cosmology was advocated by Shandarin
(1983) and Einasto et al. (1984) were the first to apply it to the study of galaxy
distributions. There are a few serious problems, however, in the straightforward
application of the percolation technique to the observed galaxy catalogs. The most
obvious one is the complicated sample geometry; as the opposite sides of a sample
361
2002 by CRC Press LLC
352
Structure statistics
Figure 10.9 A percolating cluster (for filled regions) for a P k 1 cosmological simulation. Its main features are compact clusters and filamentary bridges that connect the
clusters together. (Reproduced, with permission, from Colombi, Pogosyan, and Souradeep
2000, Phys. Rev. Let. 26, 55155518. Copyright 2000 by the American Physical Society.)
are not easy to define, percolation has to be established by following the growth
of the diameter (maximum extent) or the volume (mass) of the largest cluster. The
most serious problem, however, is the Poisson noise caused by the discreteness of
the point set. As galaxy catalogs are usually magnitude limited, their mean density drops toward the sample boundaries and we should use a distance-dependent
neighborhood radius to track structures through those regions. This can be done,
but it also amplifies the Poisson noise.
Thus, in later years, percolation analysis has been used mostly to analyze results
of cosmological numerical simulations. A thorough review is given by Klypin and
Shandarin (1993); we shall follow their presentation below. As these simulations
produce a density field on a lattice, we can carry over the results of percolation
studies from condensed matter physics. Percolation techniques have a long history
in this branch of science; a good review is given in Ziman (1979). For a density
field the role of the neighborhood radius is taken over by the critical overdensity
362
2002 by CRC Press LLC
353
level c ; we mark all lattice points of overdensity c as filled and other points
as empty and form clusters of lattice points, considering a given number of nearest
lattice points as neighbors (in three dimensions, usually six).
The structure of a typical percolating cluster for cosmological large-scale structure is rather complex, as illustrated by the three-dimensional view in Fig. 10.9.
This describes the result of a scale-free simulation with P (k ) k 1 (appropriate for scales from a few tens to a few hundreds of Mpc). We see that the network consists of more or less spherical clusters, connected by filamentary bridges.
Colombi, Pogosyan, and Souradeep (2000) show that local features of the cluster
can be classified on the basis of the three-dimensional curvature of the density
field.
The overdensity level determines the fraction of filled points (cells) p. Every
cell can be either empty with the probability 1 p, can be a member of the percoalso the infinite cluster),
lating cluster with a probability 1 (this cluster is called P
or can be a member of a finite cluster with a probability v vn(v ). Here v is the
size (number of cells) of a cluster and n(v ) is density of clusters of size v (the
multiplicity function). These quantities describe all properties of clustering on the
lattice; they have to satisfy the obvious relation
1 p + 1 +
X
v
vn(v) = 1:
Two clustering characteristics are used in practice to check for percolation. The
first is the normalized volume of the largest cluster
1 = vmax =N 3 (p pc ) ;
(10.6)
where vmax is the number of cells in the largest cluster and N 3 is the total number
of grid cells (we have assumed a cubic grid volume of N cells per side, for simplicity). The power-law relation near the percolation threshold pc is true for large
filling factors p > pc (1 = 0; p < pc ) and is predicted by condensed matter
physics. The value of pc for a Poisson lattice (randomly assigned filled and empty
cells) is pc 0:313. The exponent = 0:4 for a Poisson lattice has been found
not to differ much for cosmological simulations (Klypin and Shandarin 1993 get
0:5). At percolation the 1 (p) curve (the percolation curve) grows very
fast, but the exact percolation threshold is not easy to determine from this curve
alone. A better estimate of the percolation threshold pc can be found by fitting
(10.6) to the percolation curve.
The second characteristic is the weighted mean square size of all other (finite)
clusters
2 =
1 Pv v2 n(v)
N 2 v n(v)
jp pc j ;
where the factor 1=N 2 is chosen for normalization and the power-law relation
comes again from condensed matter theory; the exponent
1:7 is both for the
363
2002 by CRC Press LLC
354
Structure statistics
0.75
0.5
0.25
0
-3
-2
-1
0
-3
-2
-1
Figure 10.10 Percolation statistics for various model distributions. The upper panel shows
the 1 versus overdensity level curve and the lower panel shows the 2 statistics. Different
line types correspond to different N -body models. (Reproduced, with permission, from
Coles et al. 1998, Mon. Not. R. Astr. Soc., 294, 245258. Blackwell Science Ltd.)
Poisson lattice and for cosmological models. As seen from the formula above, the
2 (p) dependence peaks at p = pc .
An example of the behavior of these two statistics is shown in Fig. 10.10.
This figure shows the percolation curves for galaxy cluster density distribution in
different cosmological N -body models. There are a few differences from the
methods described above. First, the curves are shown as functions of the normalized overdensity level = = , where is the rms overdensity. Second, the
1 statistics is determined on the basis of the volume of a cluster spanning the
calculation volume, so it is zero until the start of the percolation. We see that
both statistics are rather noisy (error bars are obtained by comparing several realizations). The dynamical evolution of the topology of the large-scale structure
364
2002 by CRC Press LLC
355
0.15
0.8
0.1
0.6
0.05
0.4
0.2
-0.05
0.15
0.8
0.1
0.6
0.05
0.4
0.2
-0.05
0.15
0.8
0.1
0.6
0.05
0.4
0.2
-0.05
0
0
0.1
0.2
0.3
0.4
0.5
0.1
0.2
0.3
0.4
0.5
Figure 10.11 Percolation and genus curves for a P (k ) k 2 N -body model for various epochs. Thick solid lines describe filled regions; thick dashed lines represent voids.
Thin lines indicate percolation levels, solid lines indicate filled regions, and dashed lines
indicate voids. The dotted line shows the initial percolation level p 0:2. Error bars are
obtained by comparing several realizations and the three panels correspond to different
epochs the higher the panel, the earlier the epoch. (Reproduced, with permission, from
Sahni, Sathyaprakash, and Shandarin 1997, Astrophys. J. Lett., 476, L1L5. AAS.)
is illustrated in Fig. 10.11, where the percolation curves 1 (p) are calculated for
three different moments for several realizations of a P (k ) k 2 N -body simulation. The epochs are described by the wavenumber kNL where the dynamics just
becomes nonlinear (larger kNL correspond to smaller scales and to earlier times).
This figure also shows another advantage of the lattice approach to percolation.
While it is difficult to define voids in the case of point samples (see Section 10.8),
voids and filled regions can be treated on a lattice in the same way; voids on a lattice are defined as clusters of empty cells. Because all cosmological simulations
365
2002 by CRC Press LLC
356
Structure statistics
start from a realization of a Gaussian density field, the initial percolation properties of voids and filled regions are identical (the Gaussian distribution is symmetric). During the evolution the percolation curves start to differ topology of
filled regions becomes more filamentary during the evolution and that of empty
regions becomes more spherical.
Fig. 10.11 also shows the genus curves for the same epochs, described earlier
in this chapter. As we see, the g (p) relation is very insensitive to dynamical evolution (but it is sensitive to the form of the power spectrum, as shown by Sahni,
Sathyaprakash, and Shandarin 1997).
Thus far, the application of percolation techniques to observed samples has
suffered from Poisson noise. With the expected rapid growth of the size of future
galaxy catalogs the situation should improve; even the volume-limited samples
necessary for percolation studies will have enough galaxies to repress the influence of noise.
10.5 Minimal spanning trees
Several statistics used to find filamentary shapes in the galaxy distribution were introduced in the previous sections. Here we show a particular technique to quantify
the filamentary character of the galaxy clustering based on the minimal spanning
tree.
The minimal spanning tree (MST for short) is a graph-theoretical construct
that was introduced by Kruskal (1956) and Prim (1957). The MST of a set of N
points is the unique network of N 1 edges (each linking two points) providing
a route between any pair of points while minimizing the sum of the lengths of the
edges. We have already used the MST in Chapter 4 to provide an estimator for
the Hausdorff dimension. Here, we present this graph-theoretical construction in
more detail. The MST can be formulated as follows.
The data set is a graph G, consisting of a vertex set V (the points) and edge
set E (an edge is a straight line connecting two points, E is a subset of V V ),
each having a length or weight (in our case the Euclidean distance between
the two vertices). A sequence of edges joining vertices is a path; a closed path is
called a circuit. If there is a path between any pair of vertices the graph is called
connected. A connected graph containing no circuits is called a tree. If the tree
of a connected graph contains all the vertices of the vertex set then it is called a
spanning tree. The length of a tree is defined to be the sum of the weights of the
component edges. The MST is the spanning tree of minimal length.
The algorithm we used for calculating the MST of a given point set starts by
choosing an arbitrary point of the set and finding its nearest neighbor. These two
points and the corresponding edge form the subtree T1 . For each isolated point
(point not yet in the subtree) the identity and distance to its nearest neighbor
within the subtree is stored; by definition, this distance is called the distance of the
isolated point to the subtree. These potential MST edges are called links. The M th
366
2002 by CRC Press LLC
357
Figure 10.12 A subsample of the CfA1 redshift survey enclosed in a cubic window. There
are 153 galaxies; their MST is shown in the right panel. Several filamentary structures are
traced by the graph. (Reproduced, with permission, from Martnez et al. 1990, Astrophys.
J., 357, 5061. AAS.)
Figure 10.13 The length distribution function, normalized to the mean edge length, of
the MST corresponding to a binomial process (dashed line) and to a Voronoi model (solid
line). The distributions reflect the clustering of the underlying point patterns. (Reproduced,
with permission, from van de Weygaert 1991, Voids and the Geometry of the Large Scale
Structure, Leiden University.)
367
2002 by CRC Press LLC
358
Structure statistics
Wavelets
359
of the wavelet transforms. For simplicity, in this section we shall deal with onedimensional wavelets. The decomposition of scale performed by the wavelet analysis is made by translation and dilation of a single parent function, which is usually called the analyzing wavelet. This is a square integrable function that has to
verify the admissibility condition
Z +1 jgb(k)j2
dk < 1
0
where
gb(k) =
jk j
Z +1
1
(10.7)
e ixk g(x)dx
gp (x) = (p + 1)
i
:
(1 ix)1+p
The other example is the Marr wavelet, which is the second derivative of a Gaussian (this is a real-valued wavelet)
2
g(x) = (1 x2 )e x =2 :
The wavelet transform of a signal f (x) with respect to the analyzing function
g is
Z +1
x b
dx;
a
1
where the star denotes the complex conjugate. The parameter a > 0 is the dilation
of the length scale and b is the position. Therefore, the wavelet transform is just
the inner product hf; gab i, where
x b
1
gab (x) = p g
:
a
a
Tg (a; b) =
p1a
f (x)g
To stress the main differences between the wavelet transform and the Fourier
transform Farge (1992) has proposed considering the Fourier transform of a smooth
function f (x) with some singularities. Although the information about f (x) remains in fb(k ) = hf (x)jeikx i, the Fourier spectrum does not describe the spatial
distribution at all, i.e., it is not possible to find the singularities in the Fourier
space. On the contrary, the local character of the wavelet transform allows us to
localize the singularities; in fact, the wavelet coefficients remain smooth while
f (x) is locally smooth, and in the neighborhood of a singularity, the amplitude of
the coefficients take remarkably high values.
369
2002 by CRC Press LLC
360
Structure statistics
Figure 10.14 Two examples of wavelet functions together with their Fourier transforms.
(Reproduced, with permission, from Martnez, Paredes, and Saar 1993, Mon. Not. R. Astr.
Soc., 260, 365375. Blackwell Science Ltd.)
One interesting property of the wavelet window function is the fact that it has
zero mean
Z
g(x)dx = 0:
Wavelets
361
We shall use this property in the calculation of the Holder exponents from the
scaling of the wavelet transforms.
For example, we may ask where the singularities of the fractal measure lie, or
in other words, whether it is possible to know the spatial locations of the different
structures interweaving in the multifractal object. This goal may be reached by
means of the wavelet transform. If the fractal measure defined over IRn is represented by d(x) = f (x)dn x, f (x) being the density, the mass within a ball
centered at the point y, and radius r is just the integral of d(x) over that ball,
B (y;r)
d(x):
In IRn , the wavelet transform of the measure with respect to the analyzing function g (x) is defined by (Argoul et al. 1989),
1
Tg (a; b) = m g (a 1 r 1 (x b))d(x) a > 0; b 2 IRn ; (10.8)
a
where r represents the n-dimensional rotation operator. The exponent m is freely
chosen to get the best visual representation.
One of the most frequently used wavelet functions is the Marr wavelet (Argoul
et al. 1989; Slezak, Bijaoui, and Mars 1990):
g(z) = (n
jzj2 )e jzj2=2 :
This function is also called the Mexican hat wavelet in 2-D for obvious reasons
(see Fig. 10.15). Due to its radial symmetry, this analyzing wavelet works well
only for isotropic processes. Therefore, the rotation operator in Eq. 10.8 can be
dropped. It also should be possible to use anisotropic wavelets. The possibility
of using nonisotropic wavelets could be very interesting in the analysis of the
specific shapes introduced by the peculiar velocities in the redshift catalogs.
The wavelet transform may be considered as a mathematical microscope (Arneodo, Grasseau, and Holschneider 1988), whose magnification and position are
given by a 1 and b, respectively. The optics of this microscope should be determined by the choice of g . This is illustrated in Fig. 10.16, where the wavelet
transform of the the first CfA2 slice is shown for different values of a. In this illustration, the galaxy distribution has been considered two dimensional, ignoring
the declination.
This variation can be used as a measure of the local scaling properties of the
multifractal set due to the fact that wavelet transforms inherit the preservation of
the scaling. As a consequence of the scaling invariance of the wavelet transform,
if the measure has multifractal scaling behavior
(B (y; r)) r(y)
the wavelet transform reproduces the scale invariance in the following way
(Arneodo, Grasseau, and Holschneider 1988; Ghez and Vaienti 1989; Freysz et
371
2002 by CRC Press LLC
362
Structure statistics
Figure 10.15 The Mexican hat wavelet function. (Reproduced, with permission, from
Martnez, Paredes, and Saar 1993, Mon. Not. R. Astr. Soc., 260, 365375. Blackwell Science Ltd.)
al. 1990)
b(b) = (b) m:
The last equation implies that we can obtain the Holder exponent (b) associated
with a singularity located at b, just by taking the slope of the loglog plot of the
absolute value of the wavelet transform versus the scale a (Martnez, Paredes, and
with
Saar 1993).
Wavelets were introduced in astronomy by Slezak, Bijaoui, and Mars (1990).
These authors used the wavelet transform as a structure identifier in cluster analysis performed on Schmidt plates. It also can be used as a powerful tool in the
analysis of the subclustering in rich clusters of galaxies (Escalera and Mazure,
1992) and for detecting voids and high-density regions in the CfA slice (Slezak,
de Lapparent, and Bijaoui 1993), as shown in Fig. 10.17.
10.7 Cluster-finding algorithms
The definition of a cluster in terms of a given distribution of galaxies seen in
projection on the sky is a complex process, involving both the nature of the original survey and the selection criteria. Certainly, in the past, the cluster identification
methods were subjective and difficult to quantify, yet even now with automatically
372
2002 by CRC Press LLC
Cluster-finding algorithms
363
Figure 10.16 The wavelet transform acting on the first CfA2 slice (see Fig. 8.4). We see
how different values of the magnification parameter reveal different structures of the galaxy
distribution. (Reproduced, with permission, from Martnez, Paredes, and Saar 1993, Mon.
Not. R. Astr. Soc., 260, 365375. Blackwell Science Ltd.)
scanned survey plates and with computer algorithms being used to find and classify clusters, the situation does not appear to be that much simpler.
What is clear is that we can identify isolated galaxies simply by looking to see
how many neighbors galaxies have within a specified radius. Successive removal
of galaxies that are isolated according to this criterion from a sample leaves a set
of galaxies that are clustered, and then we can select subsets of those that define
clusters having specific properties. The clusters found by this method depend on
the choice of a length scale and a density contrast.
An alternative and quite popular strategy is to identify clusters of points by a
friends-of-friends algorithm. Two galaxies are within the same cluster if they
are within a specified distance of each other. This is closely related to the identification of clusters by graphical methods such as the MST construct. The clusters
found by this method depend on the choice of a length scale denoting how close
a galaxy must be to another to be attached to the same cluster.
373
2002 by CRC Press LLC
364
Structure statistics
Figure 10.17 The wavelet transform is used here to detect the voids in the galaxy distribution of the CfA2 slice, after removal of the Coma cluster. (Reproduced, with permission,
from Slezak, de Lapparent, and Bijaoui 1993, Astrophys. J., 409, 517529. AAS.)
10.7.1 MST
The minimal spanning tree (MST) introduced in Section 10.5 may be used as a
base to find clusters. Fig. 10.18 shows the MST associated with clustered point
process within a box of 100h 1 Mpc sides. Here we describe how the MST
might be used to select clusters. Given a length scale rs , if we remove the edges
with length greater than rs , different disconnected subsets still remain linked with
edges smaller than rs . We can consider all these subsets as clusters associated with
the length scale rs . Obviously, as rs decreases, the number of clusters increases,
while their size gets smaller. This effect is clearly illustrated in Fig. 10.19, which
shows the clusters selected by means of this pruning method for the value of
rs = 2 h 1 Mpc.
10.7.2 Modified friends-of-friends
In Section 10.4 we explained the friends-of-friends algorithm introduced by Press
and Davis (1982). A modification of this technique was introduced by Croft and
Efstathiou (1994) to select clusters from their simulation N-body simulations. Let
us illustrate how this process works on our clustered point process. For each point
of the process, we count how many neighbors lie within a distance rs , and we
consider all these balls (as many as points) as the first generation of clusters. We
find the center of mass of each cluster and delete any cluster within rs of a cluster
of larger mass. The process continues by counting the number of particles within
balls of radius rs placed in the centers of mass of the surviving clusters. These
374
2002 by CRC Press LLC
Cluster-finding algorithms
365
Figure 10.18 A 2-D realization of clustered point process: (a) the point set; (b) the corresponding MST. (Reproduced, with permission, from Paredes, Jones, and Martnez 1995,
Mon. Not. R. Astr. Soc., 276, 11161130. Blackwell Science Ltd.)
are the clusters of the second generation. Again, we compute the new centers
of mass and delete in the same way the overlap clusters (we consider that two
spheres have enough overlap if the distance between the centers is smaller than
their radius). The process is repeated several times. Croft and Efstathiou (1994)
start the process from the percolation centers, instead of from all the points, as
tested here.
Fig. 10.20 shows how the clusters are selected using this algorithm for values
of rs = 1 and rs = 2 h 1 Mpc.
375
2002 by CRC Press LLC
366
Structure statistics
Figure 10.19 Illustration of pruning the MST to select clusters from a clustered point process with length scale rs = 2 h 1 Mpc. We can see the surviving connected branches of
the tree and the centers of mass of the corresponding clusters. (Reproduced, with permission, from Paredes, Jones, and Martnez 1995, Mon. Not. R. Astr. Soc., 276, 11161130.
Blackwell Science Ltd.)
Figure 10.20 Clusters selected using the modified friends-of-friends algorithm. In the left
panel, rs = 1 h 1 Mpc, and in the right panel rs = 2 h 1 Mpc. (Reproduced, with
permission, from Paredes, Jones, and Martnez 1995, Mon. Not. R. Astr. Soc., 276, 1116
1130. Blackwell Science Ltd.)
376
2002 by CRC Press LLC
Cluster-finding algorithms
367
Figure 10.21 Clusters selected using the algorithm based on the wavelet transform for the
value of rs = 1 h 1 Mpc (a = 0:93). The bottom plot shows the isocontours of the wavelet
coefficients. (Reproduced, with permission, from Paredes, Jones, and Martnez 1995, Mon.
Not. R. Astr. Soc., 276, 11161130. Blackwell Science Ltd.)
10.7.3 Wavelets
Wavelet transforms can be used to select clusters from the point process. We have
seen how the wavelet transform acts as a mathematical microscope in which b
is the location and a 1 is the magnification. Therefore, changing the scale a, the
wavelet transforms reveal different structures as a function of their size.
The bottom panel of Fig. 10.21 shows the isocontours of the wavelet transform
applied to our clustered set for a = 0:93. If a is increased the small structures are
erased. For a given value of a, we will consider, as the centers of the clusters, the
local maxima of the function Tg (a; b). After erasing overlapping clusters as in the
377
2002 by CRC Press LLC
368
Structure statistics
previous method, we end up with a set of positions for the selected clusters. The
top panel of Fig. 10.21 shows how this method selects clusters in the clustering
model for rs = 1.
A comparison of the efficiency of the previous methods is presented in Paredes,
Jones, and Martnez (1993). For finding rich clusters the three methods provide
identically good results: they are equally efficient. The differences appear when
the methods are used to find small groups formed by pairs, triplets, and so on. The
dependence on the scale length rs is then crucial.
The MST method is very sensitive to the chosen value of the pruning length rs .
This is because, if rs is large, different clusters percolate and are considered as a
single structure. The wavelet method works well for finding clusters, although it
does miss some that are found by the modified friends-of-friends algorithm. One
of the most promising techniques that can be exploited for finding clusters from
the homogeneous and isotropic matter distribution was recently introduced by
Sanz, Herranz, and Martnez-Gonzalez (2001). These authors present a method
with very good performance based on the selection of optimal pseudofilters once
the density profile of the sources to be detected has been assumed. The method
provides an unbiased and efficient estimator of the amplitude of the source.
The quantity of new cosmological data that will become available in the next
years (2df, SDSS) necessitates the development of more efficient algorithms for
finding clusters in multidimensional databases. This effort is being made as part
of a huge Computational AstroStatistics collaboration (Nichol et al. 2000). The
algorithms developed by this collaboration will be part of the statistical analysis
tools of the Virtual Observatory (see Virtual).
10.8 Void statistics
Voids are easy to see in the galaxy distributions. To study their properties, we
need an operational definition of a void. Surprisingly, this concept has been rather
difficult to define. In the first study of voids Einasto, Einasto, and Gramann (1989)
used the definition of a void as an empty sphere of a maximum radius. This can
be made more precise, defining the distance field (Aikio and Mahonen 1998) by
D(x) = min
fjx xi jg;
i
(10.9)
where xi are the coordinates of sample points (galaxies). The local maxima of
D(x) are then the centers of spherical voids and the values of D at these points
are the void sizes.
Of course, this void definition can be used to compare observational samples
with numerical models, but it does not correspond to our intuitive notion of a void.
To remedy this, a number of algorithms were later proposed. We describe two that
give the closest result to the voids we detect visually.
The first algorithm is the void finder algorithm of El-Ad and Piran (1997).
Its logic is probably very close to the human void recognition mechanism. We
378
2002 by CRC Press LLC
Void statistics
369
Voronoi cells
galaxies
walls
d = 0.816
d = 0.680
d = 0.510
d = 0.410
d = 0.356
Figure 10.22 The void finder applied to a Voronoi tessellation. The first two panels show
the original tessellation and the galaxies generated at cell walls. The third panel shows the
result of the work of the wall builder. Subsequent panels show successive stages of the void
finder at work. (Reproduced, with permission, El-Ad and Piran 1997, Astrophys. J., 491,
421435. AAS.)
demonstrate how it works in Fig. 10.22. Basically, this algorithm determines voids
as unions of empty spheres of different radii. The problem with such a natural
definition is that it is difficult to prohibit percolation of voids. We know that both
in numerical simulations and in observations the empty regions percolate, but we
perceive neighboring voids as different entities. For this reason their algorithm
starts with the wall builder that separates galaxies into two classes, the wall
galaxies that delimit voids and field galaxies that do not. Wall galaxies are
defined as galaxies that have at least n other wall galaxies at a distance l around
it. The values of n and l are free parameters of the algorithm; El-Ad and Piran
choose n = 3 and l, which is related to the density of the sample. The wall
builder has done its work by panel 3 (walls) of Fig. 10.22.
After the walls have been defined, the void finder starts fitting empty spheres in
the voids. This process works on a grid in sample space and proceeds in steps of
different sphere sizes. First, unions of spheres of slightly different radii are found
that fit into the voids. This produces the fourth panel (d = 0:816) in Fig. 10.22.
The grid points that have been assigned to a void will not be reassigned in the
future. If older voids exist, spheres of the same radii are used to add extra grid
points to them. Then the radius is decremented and the procedure is repeated until
379
2002 by CRC Press LLC
370
Structure statistics
Figure 10.23 Voids in the IRAS 1.2-Jy redshift survey. There are 24 separate voids. The
zone of avoidance is seen in the center of the figure; the Great Attractor is shown as space
devoid of voids at the left; and the Perseus-Pisces supercluster is shown at lower right.
(Reproduced, with permission, from El-Ad, Piran, and da Costa 1997, Mon. Not. R. Astr.
Soc., 287, 790798. Blackwell Science Ltd.)
371
simpler (Aikio and Mahonen 1998). This starts from the distance field D(x)
(10.9), defined on a grid. Starting now from a grid point xi we set up a path
toward the fastest growth of D(x) until the path ends at a local maximum Mj .
Then the point xi and all points on the path are assigned to the subvoid vj with
the center at Mj . In this way, all points are assigned to subvoids in one pass of the
algorithm (and if a path reaches an assigned point before a maximum, the path
can be assigned at once).
Second, subvoids are collected together into voids. If the centers of two subvoids M1 and M2 are closer than maxfD(M1 ); D(M2 )g, the subvoids are joined
together. Using the same criterion, other subvoids can be joined to the large void
in the friends-of-friends manner, forming a subvoid cluster. This completes the
algorithm all empty cells are assigned to voids.
The algorithm is fast, robust, and versatile. It has one free parameter cells
closer than a limiting distance Dmin to a point can be left free. Cells belonging to
border voids can be left unassigned, too; this is probably the best strategy for an
observed sample. The algorithm can also be used for different density levels on a
grid, defining filled and empty regions by the critical density level.
The voids that the algorithm produces are similar to our visual impression and
to those produced by the void finder.
372
Structure statistics
Figure 10.24 The distribution of 319 clusters in 25 very rich superclusters with at least 8
members in the southern galactic hemisphere (in supergalactic coordinates). The grid with
the step size 120 h 1 Mpc corresponds approximately to distances between high-density
regions across voids. (Reproduced, with permission, from Einasto et al. 1997b, Nature,
385, 139141. Macmillan Publishers Ltd.)
xP = x=P
bx=P c;
where the floor function bxc is defined to be the largest integer smaller than or
equal to x. If there is a (cubic) periodicity of period P in the distribution of the
sample points, it will show up as an inhomogeneous density distribution in the
phase cube CP = (0 (xP )i 1; i = 1; 2; 3).
If the folding method is applied for searching periodicities in astronomical time
series, different variants of the analysis of variance are used as test statistics. A
similar statistic that quantifies the notion of inhomogeneous density in the multidimensional case is the total variance of density in the phase cube. This statistic
(regularity) was proposed by J. Pelt and analyzed and applied to the cluster distribution by Toomet et al. (1999).
382
2002 by CRC Press LLC
373
Figure 10.25 Planar folding. The sample space (upper panel) is divided into smaller trial
squares (dotted lines) and all these squares are stacked together (lower panels). If the size
of the trial squares is close to the real period, there will be a clear density enhancement in
the stacked distribution (lower left panel); otherwise, the point distribution will be almost
random (lower right panel).
(P ) =
1
n2 (x ) d3 xP ;
n 2P CP P P
where nP (xP ) is the number density in the phase cube (the phase density) and
n P is the mean phase density. We can write this integral as
where
nP d xP =
2
CP
CP
nP dN;
is the number of points in the sample. Because the phase space has a
383
374
Structure statistics
1.3
1.25
1.2
1.15
1.1
1.05
1
0.95
0.9
50
100
150
200
250
d [h-1Mpc]
Figure 10.26 The solid line shows the regularity periodogram for the sample of galaxy
clusters in rich superclusters. Dotted lines show the 90 and 99% confidence regions.
unit volume, n
P
distribution as
(P ) =
N X
N
X
1
K (xi ; xj );
N (N 1) i j
(10.10)
where xi are the phase coordinates (we have dropped the index P here to avoid
clutter), and we substitute N (N 1) for N 2 to avoid bias. The kernel function
K (xi ; xj ) is used to estimate the phase density at sample points:
nP (xi ) =
N
X
j
K (xi ; xj ):
(10.11)
The kernels we can use here are the usual density kernels (cubic, Gaussian, Epanechikov) described in Chapter 3.
We usually search for small-amplitude signals, when the population distribution is almost Poissonian. Toomet et al. (1999) show that the sample distribution
of the statistic (10.10) for a Poisson population can be approximated by the 2
distribution, where the number of degrees of freedom = 1=V" and V" is the
phase volume of the density kernel. The mean value of the statistic is E fg = 1
for all N , and its sample variance is
Var =
2
:
N2
The smallness of this variance allows detection of very weak cubic signals. Toomet
et al. (1999) recommend choosing a constant width of the density kernel " in phase
384
2002 by CRC Press LLC
375
1
00
1 0
Figure 10.27 Phase cube for the best period 130 h 1 Mpc for galaxy clusters in rich superclusters in the southern galactic hemisphere. A projection of their spatial distribution
is shown in Fig. 10.24.
space for all trial periods P ; this keeps the sample variance of the test statistic
constant. Another recommendation is to discard in the density sum (10.11) those
points that are close to the point i in data space; this eliminates the effect of local
correlations. The weighting caused by the variable mean density of real samples
(the selection function) can be taken into account by normalizing the test statistic
not by 1=N 2 , as in (10.10), but by the value of the test statistic for a binomial point
process of a much larger total number of points, but with the same geometry and
selection function as the real sample. Because this test is anisotropic, the search
space for the periodogram maxima is four dimensional: periodograms depend on
the value of the trial period and on the three Euler angles, which are needed to
specify the orientation of a cube in space.
As an application of this method to the real samples, Fig. 10.26 shows the
regularity periodogram for the full cluster sample in high-density regions (the
southern galactic subcone of this sample is shown in Fig. 10.24). A periodogram is
defined as the dependence of a test statistic on the trial period; thus, the regularity
periodogram is the function (P ).
Fig. 10.26 shows also two acceptance regions (the inner region for the 90%
acceptance level, the outer region for the 99% acceptance level) for a binomial
point process with the same number of points as the number of clusters, and with
the same sample geometry and selection function. Comparison of these regions
385
2002 by CRC Press LLC
376
Structure statistics
with the sample periodogram shows that some of the peaks are significant with a
level 1%. The most significant period around 130150 h 1 Mpc is clearly
seen, but there also indications of a harmonic at 6080 h 1 Mpc and of a subharmonic at 240h 1 Mpc, as it should be in the case of a cubic arrangement.
The specific significance levels could be changed by accounting for the multiple
comparison problems, but for the present case the phase cube was oriented along
the supergalactic Cartesian coordinate axis without previous directional search,
so the correction is not too large.
Fig. 10.27 shows the spatial distribution of points in the phase cube for the best
period for the half-sample shown in Fig. 10.24.
This phase cube shows that the cubic regularity exists; the density distribution
in the phase cube is highly inhomogeneous. The cubic signal is mainly due to a
couple of rich point clusters plus several filamentary features; such a configuration
can be obtained by folding a lattice populated by points with higher concentration
at corners.
We note that the existence of such a regularity is still controversial. The significance of the test would be smaller if we compared the observed sample with
a Poisson cluster process, and the period is too close to the size of the cluster
samples (about one third of the depth of a subcone). Finally, there is no known
theoretical explanation for a cubic arrangement in the large-scale structure.
This all could mean that what we observe is a statistical fluke. The method,
however, exists.
386
2002 by CRC Press LLC
APPENDIX A
Coordinate transformations
A.1 Introduction
Positions of objects in the sky are described in various spherical coordinate systems. When studying galaxy catalogs, we frequently need to transform from one
system to another, and these formulae can be difficult to find. To make this task
easier, we describe here the main coordinate transformations and give the necessary formulae. We also list a few of the more popular map projections that are
used to visualize the distribution of objects in the sky.
Observing the stars and other objects in the sky, we can imagine that they lie on
a sphere (the celestial sphere). The projection of the Earths rotational axis intersects the celestial sphere at the north celestial pole (NCP) and the south celestial
pole (SCP). Likewise, projections of the terrestrial meridians and parallels onto
the sky form the celestial meridians and parallels.
A.2 The equatorial system
The equatorial system is the basic system used to give the positions of objects
in catalogs. The projection of the equator of the Earth onto the celestial sphere
is the celestial equator. This great circle is the basis of the equatorial system of
astronomical coordinates used to assign positions to the stars and other celestial
objects.
The declination of a star, , is the angular distance between the star and the
celestial equator. The declination takes values from 0 to 90 for objects in the
northern celestial hemisphere and from 0 to 90 for objects in the southern
hemisphere. This spherical coordinate plays the role of the terrestrial latitude.
To define the second coordinate, which should be similar to the terrestrial longitude, we need to fix the zero meridian, similar to the role played by the Greenwich
meridian. This is done using another great circle on the celestial sphere, called the
ecliptic. The ecliptic is the great circle described by the Suns trajectory as a consequence of the orbital motion of the Earth around the Sun in the course of a year.
The celestial equator and the ecliptic form an angle of = 23 260 2900 ' 23:5 .
Therefore, these two great circles cross each other at two points. One of this
points, the Vernal equinox, is the one in which the Sun lies when it crosses the
equator from the southern hemisphere to the northern hemisphere, at the spring
equinox. This is the direction used for defining the zero point of the second
387
2002 by CRC Press LLC
378
Coordinate transformations
388
2002 by CRC Press LLC
Coordinate transformations
379
Figure A.1 Relation between the galactic and the equatorial coordinates.
coordinates (SGX; SGY; SGZ ) are also frequently used, with the supergalactic
z -axis directed toward the NSP, the x-axis toward the origin, and the y-axis perpendicular to the first two.
A.4 Coordinate transformations
By means of simple relations from spherical trigonometry, it is straightforward
to obtain the equations to transform coordinates from one system to another. We
follow here another approach used by the Groningen Image Processing System
(see its home page at Gipsy). It uses Cartesian 3 3 transformation matrices to
convert coordinates. Let us suppose that we want to change the coordinates (; )
of an object to a different system (0 ; 0 ), where is the longitude and is the
latitude. The unit vector pointing to the direction of (; ) has coordinates
r1
r2
r3
= cos cos ;
= sin cos ;
= sin :
(A.1)
s Tr
0
0
=
=
380
Coordinate transformations
The matrices for useful coordinate transformations are the following (taken
from the tables in skyco.c in the Gipsy source code):
From equatorial 1950.0 to equatorial 2000.0
0
0:9999257453
= @ 0:0111761178
0:0048578157
0:0111761178
0:9999375449
0:0000271444
0:0048578157
0:0000271491 A :
0:9999882004
0:0548808010
= @ 0:4941079543
0:8676666568
0:8734368042
0:4448322550
0:1980717391
0:4838349376
0:7469816560 A :
0:4559848231
0:3751891698
= @ 0:8982988298
0:2286750954
0:3408758302
0:0957026824
0:9352243929
0:8619957978
0:4288358766 A :
0:2703017493
T=@
0:0548808010
0:8734368042
0:4838349376
0:4941079543
0:4448322550
0:7469816560
0:8676666568
0:1980717391 A :
0:4559848231
0:7353878609
= @ 0:0745961752
0:6735281025
0:6776464374
0:0809524239
0:7309186075
0:0000000002
0:9939225904 A :
0:1100812618
0:3751891698
= @ 0:3408758302
0:8619957978
0:8982988298
0:0957026824
0:4288358766
0:2286750954
0:9352243929 A :
0:2703017493
0:7353878609
= @ 0:6776464374
0:0000000002
0:0745961752
0:0809524239
0:9939225904
0:6735281025
0:7309186075 A :
0:1100812618
Sky projections
381
90
-180
180
-90
Figure A.2 The HammerAitoff full-sky projection. The coordinates are given in degrees,
and the parallels and meridians are drawn with a 30 step.
The formulae below are taken mostly from the Groningen Image Processing System (Gipsy) manual. We designate the sky coordinates as the longitude and the
latitude , and the Cartesian map coordinates as x and y .
The first class of projections are those used for full-sky maps. The most popular
all-sky projection is the HammerAitoff projection. It is an equal-area elliptical
projection (the lines of both constant longitude and latitude are ellipsoidal arcs)
with minimal distortions:
p cos sin 2
x = 2R 2 p
;
1 + cos cos 2
p
sin
:
y = R 2p
1 + cos cos 2
Here and below R is the scaling factor. Fig. A.2 shows the coordinate grid for the
HammerAitoff projection. All equal-area projections distort angles.
A simpler equal-area projection is the Mollweide projection. It is a pseudocylindrical projection where the lines of constant latitude (parallels) are represented by straight lines in the map:
p
cos ;
p
= 2R 2
2 sin
382
Coordinate transformations
90
90
-180
180 -180
180
-90
-90
Figure A.3 The Mollweide (left panel) and Sanson (right panel) full-sky projections. The
coordinates are given in degrees, and the parallels and meridians are drawn with a 30
step.
The Mollweide projection has more severe distortions than the HammerAitoff
projection.
A simpler projection is the Sanson sinusoidal projection, a pseudo-cylindrical
equal-area projection where the parallels are equally spaced:
x
y
=
=
R cos ;
R:
This projection is more distorted than both previous projections. Fig. A.3 shows
both the Mollweide projection (left panel) and the Sanson projection (right panel).
The above formulae have been written for the case when the map pole (the
center of the map x = y = 0) corresponds to the point = = 0 in the sky. The
map pole can be easily shifted to any other point on the celestial sphere.
For partial sky maps azimuthal projections are used, where the points on the
sphere are projected to the map plane that is tangential to the sphere at a specific
pole, using rays extending from the projection pole.
For small regions of the sky the gnomonic projection is used:
cos sin( 0 )
sin sin 0 + cos cos 0 cos(
sin cos 0 cos sin 0 cos(
R
sin sin 0 + cos cos 0 cos(
;
0 )
0 )
:
0 )
Because azimuthal projections are used for isolated patches of the sky, we explicitly include the coordinates 0 ; 0 of the map pole in the formulae. This projection
distorts both areas and angles.
392
2002 by CRC Press LLC
Sky projections
383
45
45
-45
45
-45
45
-45
-45
Figure A.4 The gnomonic (left panel) and orthographic (right panel) partial sky projections. The coordinates are given in degrees, and the parallels and meridians are drawn
with a 15 step.
x
y
=
=
R cos sin( 0 );
R sin cos 0 cos sin 0 cos( 0 ):
This projection distorts both areas and angles. Fig. A.4 shows the gnomonic and
orthographic projections of a 90 90 patch of the sky.
For the stereographic projection the projection pole is moved to the point on
the sphere opposite to the map pole. This gives a conformal map, where all angles are locally preserved (the areas are distorted). The stereographic projection
is frequently used to represent whole hemispheres:
cos sin( 0 )
1 + sin sin 0 + cos cos 0 cos(
sin cos 0 cos sin 0 cos(
2R
1 + sin sin 0 + cos cos 0 cos(
2R
;
0 )
0 )
:
0 )
Fig. A.5 shows the stereographic projection for two map poles, for the usual =
= 0 pole in the left panel and for the north pole of the sky ( = 90 ) in the right
panel.
Another projection that is frequently used for projections of whole hemispheres
is the Lambert azimuthal equal-area projection. The map pole is usually chosen
to coincide with the celestial north or south pole, and the formulae for this case
are:
x
y
=
=
R 2
p
R 2
2j sin j cos ;
2j sin j sin :
Fig. A.6 shows the Lambert azimuthal equal-area projection for the northern sky.
393
2002 by CRC Press LLC
384
Coordinate transformations
0
90
30
60
-90
90
90
-90
Figure A.5 The stereographic projection. The left panel illustrates the case where the map
pole is at the origin of the spherical coordinates ( = = 0). The right panel shows
the projection when the map pole coincides with the celestial north pole ( = 90 ). The
latitudes of the parallels are given in degrees, and the meridians are drawn with a 30
step.
0
30
60
90
Figure A.6 The Lambert azimuthal equal-area projection. The map pole coincides with
the celestial north pole ( = 90 ). The latitudes of the parallels are given in degrees, and
the meridians are drawn with a 30 step.
394
2002 by CRC Press LLC
APPENDIX B
386
B.3 Estimation
We are usually interested in finding information about the properties of a whole
parent population, but we have only the data in our sample. The procedure that
allows us to do that is called estimation.
A rule or a method of estimating a parameter of the parent population is called
an estimator. It is usually given as a function of observables (sample values). A
value of an estimator, found for a given sample, is called the estimate.
can be used to estimate the population mean :
Example: The sample mean x
N
1 X
x;
(B.1)
N i=1 i
b denotes, by convention,
where N is the size of the sample fx1 ; : : : ; xN g and
the estimator of .
b = x =
Any function of observables alone is called a statistic. Estimators are also statistics, but the class of statistics is larger.
Example: The F -statistic is used to compare the variances of two populations,
using two samples:
V1
;
V2
N
X
1 i=1
(xi
x)2
Properties of estimators
387
then the function f () as a parameter of for the fixed sample values is called the
likelihood function or the likelihood of the sample. It is usually designated as L,
and it is the probability density of selection of that particular sample for a given
. For simplicity, we shall consider below mainly the case of one parameter; in
the case of a larger number of parameters k we can always substitute with a
parameter set (1 ; : : : ; k ).
Example: The likelihood function for a random sample from the population with
the probability density f (x; ) is
L() =
Given the sampling distribution
value E ftg of any statistic t by
E ftg =
f (xi ; ):
t dF (x1 ; : : : ; xn ):
The expectation value is the mean value of a statistic over all possible samples
of the same size. Note that a more extended notation in statistics for (B.3) is ET
with capital T for the estimator.
B.4 Properties of estimators
The goodness of an estimator is judged on the basis of several desirable properties.
An estimator should be consistent, meaning that it should converge in probability
to the true value of the population parameter when the sample size increases.
More exactly, any estimator tn for a parameter , found from a sample of size
n, is consistent, if for any positive small " and there is some N such that the
probability that
jtn j > "
is greater than 1 for all n > N . Consistency can also be defined for other
types of convergence.
Example: The sample mean (B.1) is a consistent estimator for the population
mean, if the distribution is Gaussian, as seen from the previous example.
Example: The sample mean for a sample drawn from a population with the
Cauchy distribution
1
dx
dF (x) =
(B.2)
1 + (x )2
has the same distribution (B.2) as the parent population, which does not depend
on the sample size. Thus, in this case the sample mean is not a consistent
estimator for the parameter . Note that the mean of the Cauchy distribution is
undefined, while higher moments diverge.
397
2002 by CRC Press LLC
388
A consistent estimator converges to the parameter value for large sample sizes.
An estimator t of is called unbiased if provides, on average, a correct estimate
of the value of the parameter for any sample size.
E ftg = ;
(B.3)
If the above relation (B.3) is not satisfied, the difference b(t), defined by
b(t) = E ftg ;
is called the bias. If we can calculate the bias for an estimator, we can correct
that estimator to make it unbiased. If there exists an unbiased estimator for a
parameter, the parameter is called U-estimable.
Example: Consider a sample of size N from a population with mean and vari is
ance 2 . The expectation value of the sample mean x
E fxg = E
1 X
N i
xi
1 X
N i
E fxi g =
1 X
N i
= :
E fV g
=
=
E
8
<
1 X
N i
(xi
x)2
N 1X 2
E
xi
: N2
i
N
E fx2 g
=E
8
<
1 X
:N
1 XX
xi
9
=
1 X
N i
xi
!2 9
=
;
xx
N 2 i j 6=i i j ;
N 1 2 N 1 2
=
:
N
N
Vb
1 i
(xi
x)2 :
For an unbiased estimator, we should not expect that half of the estimates are,
on average, above and half below, because the central measure used in the definition of unbiasedness in relation (B.3) is the mean and not the median.
An estimator should also be given as a function of sufficient statistics. A statistic
t is sufficient for a parameter if it contains all the relevant information in the
sample about and no other statistic can add anything to that. This condition is
more useful if written for the likelihood function L:
(B.4)
Properties of estimators
389
where the function g (t; ) depends only on t and . If the number of population
parameters is larger than one (say, k ), we have to consider instead of a single
statistic t a set of sufficient statistics (t1 ; : : : ; tm ). It is easily seen from (B.4)
above that the observations themselves always constitute a set of sufficient statistics. Examples of sufficient statistics are given below.
An estimator with a large sample variance is not a good estimator. Therefore, a
desirable property for an estimator is to present the smallest possible variance. In
practical cases, a compromise has to be reached between unbiasedness and minimum variance. For example, if we select a freely chosen constant as an estimator,
its variance is trivially zero, but it might be strongly biased (Meyer 1975). For
some parameters it is possible to find the minimum variance unbiased estimator
(MVUE). The MVUE does not have to exist but it does exist under certain regularity conditions (see Lehmann and Casella 1998). For an unbiased estimator t
of the parameter the minimum variance varmin (t) is given by the CramerRao
inequality
var(t) varmin = E
8
>
<
>
:
9
>
=
8
>
<
9
>
=
:
= E
2
>
@ ln L 2 >
;
;
: @ ln L >
@
@2
(B.5)
If one estimator has a smaller sample variance than another, it is said to be more
efficient. If the MVUE exists, the efficiency of an estimator t is defined as the ratio
of the minimum variance varmin to its sample variance var(t):
varmin
:
eff(t) =
var(t)
If the sample variance of an estimator coincides (for large samples) with the
MVUE, the estimator is said to be efficient.
Example: The sample mean is an efficient estimator for the population mean.
As the MVUE are, in general, difficult to find, another way to measure the
performance of an estimator is with its mean square error (MSE), which equals
the square of the bias plus the variance: b(t)2 + var(t).
A general efficient estimator of is the maximum likelihood estimator (MLE),
defined as the value b that maximizes the likelihood function:
@L
= 0;
@ =b
@2L
< 0:
@2 =b
The efficiency of the ML estimator allows the use of (B.5) for estimating its variance when the sample is large. Maximum likelihood estimators are invariant under
parameter transformations (minimizing the likelihood function with a respect of
a function g () will give the same result as minimizing it with respect of ), but
they can be biased. However, ML estimators are consistent, so they are asymptotically unbiased. In addition, if a sufficient statistic exists, the ML estimator is a
function of it. For a review of ML methods we recommend Le Cam (1990).
399
2002 by CRC Press LLC
390
Example: The likelihood function L for a sample drawn from a univariate (onedimensional) Gaussian population is
1
exp
L=
(2 )N=2 N
"
1 X
2 2 i
(xi
)2 ;
where N is the size of the sample, and and 2 are the population mean and
variance, respectively. Instead of maximizing the likelihood function L it is
usually easier to maximize the log-likelihood function
1 X
L = ln L =
Minimizing
2 2 i
)2
(xi
ln(2 )
ln +
2
(xi
b3
i
b)2
(xi
X
b)
= 0;
= 0:
b
i
b =
b2 =
1 X
N i
xi = x;
1 X
(x
N i i
x)2 :
xi f (xi ; ) 2
:
i
i
This estimator is called the least squares estimator and it is frequently used as an
independent postulate, without looking for justification by the postulate of maximum likelihood.
The last desirable property of an estimator is robustness. This property means
that the estimate does not depend too much on small variations of the underlying
probability distribution or the presence of outliers in the data (such as misinterpreted observations).
400
2002 by CRC Press LLC
391
Most astronomers have been taught probability in the sense of a limiting frequency, but the above quotation should not surprise us probability is the point
where mathematics meets the real world.
Barlow (1989) lists three different notions of probability:
1. Frequentist probability, where the probability of a event is determined by the
ratio of the number of times M the event occurred in N experiments, p =
M=N , when N ! 1. This definition requires that the experiment should be
repeatable. Thus, probability is not the property of a particular event, but the
property of the ensemble or parent population and the experiment. This is the
orthodox definition of probability.
2. Objective probability, or propensity, which belongs to an object the probability that a good die, when thrown, will show the numbers 1 to 6 with equal
probability, can be considered the property of that die.
3. Subjective probability is a degree of belief, allowing us to assign probability
to our everyday predictions or to our theories. The methods using that notion
of probability are grouped under the term Bayesian statistics.
B.5.2 Bayesian methods
Bayesian statistics starts with prior distributions that are updated in light of new
data by means of Bayes theorem. The theorem states that the posterior probability
401
2002 by CRC Press LLC
392
density of a statistical model (the probability we can assign to our model after an
experiment or observations) is:
f (jx1 ; : : : ; xn ; I ) = f (jI )
f (x1 ; : : : ; xn j; I )
;
f (x1 ; : : : ; xn jI )
(B.6)
where f (jI ) denotes the conditional distribution of the parameter , given a value
of I (similarly for other distributions), and I denotes the prior information we
have about the universe (and about similar data sets, in particular). The first factor
on the right-hand side of (B.6) is the Bayesian prior (or simply prior), which
includes our prior knowledge about the values of the parameter . The numerator
is the likelihood function (note that it can also depend on the prior information, in
principle) and the denominator is the normalization constant (global likelihood).
The latter frequently can be difficult to find, because it gives the probability of
data conditional on the knowledge of all possible values of the parameters. The
Bayesian formula (B.6) shows how an observation (described by the likelihood
function) can modify our previous knowledge (the prior) of a theory.
Whereas in the frequentist approach the model (population) parameters are
considered as fixed, and we calculate probabilities for an observed sample (we
did that already in several places above), the Bayesian approach considers the
observed sample as fixed, an ultimate truth, and assigns probabilities to models
(sample parameters).
The problems usually associated with the Bayesian approach, the need to define
the prior distribution f (jI ) and difficulties in calculating the global likelihood,
really stress the need to define exactly the statistical model we compare with observations. Using the frequentist approach, we usually define a general statistic
and study its properties for different classes of statistical models, which takes at
least the same amount of work. If we do not have any prior information, a noninformative prior that does not depend on the model, for example, a flat prior, is
used. In this case the Bayesian methodology gives the maximum posterior probability for the maximum likelihood solution.
B.5.3 Confidence intervals
Different notions of probability cause different (and conflicting) understandings
of interval estimation.
For the frequentist approach, the estimators are random variables, and there is
only one value of the population parameter, although unknown. There are only
two values for the probability that the estimate t is larger than the parameter , 0
or 1. But because the estimator t is a random variable, it may take different values
for different samples, and we can find an interval of estimates (t ; t+ ) that will
include the value of the parameter with a given confidence . Such an interval
is called a confidence interval. We can choose many different confidence intervals
for a given confidence level . The most popular choices are:
402
2002 by CRC Press LLC
393
1. The central interval: the probabilities above and below the interval are equal.
2. The symmetric interval: t+
= t .
3. The shortest interval: for a given , t+ t is minimal.
The assertion that the parameter lies in a confidence interval of confidence level
will be true, on average, in a proportion of the samples of the same size.
Example: For a Gaussian sampling distribution, all three types of intervals coinb
b;
b+
b ] corresponds to the 68% confidence
cide; the confidence interval [
b 2
b;
b + 2
b ] corresponds to the 95.4% confidence level;
level; the interval [
b
2:58
b;
b + 2:58
b ] corresponds to the 99% confidence
and the interval [
level.
We can also define the one-tailed limits, the upper and lower limits. At a given
confidence level and upper limit t+ , the assertion that < t+ will be true, on
average, for a proportion of similar samples.
Example: For a Gaussian sampling distribution, the upper limit for the 68% conb + 0:47
b ; the upper limit for the 95.4% confidence level is
fidence level is
b + 1:68b; and the upper limit for the 99% confidence level is b + 2:33b.
Compare this with the previous example.
We sometimes read about fiducial intervals (Wang 2000). In the fiducial approach we consider the likelihood function for a sample as the probability density
for the (random) population parameters, given the sample values. In the fiducial
approach the assertion that the parameter lies in a fiducial interval of the level
is true with the probability for a given sample. The fiducial approach has been
shown to meet with paradoxes, especially in multivariate cases (see, e.g., Kendall
and Stuart 1967); the Bayesian approach, which also considers population parameters as random, is free of those paradoxes.
In the Bayesian approach, confidence values for parameter intervals (Bayesian
intervals) are obtained by integrals of the posterior probability over the intervals.
When an interval contains all points with the posterior probability density above
a certain level, it is called the highest posterior density interval. When the posterior probability density is found, the intervals are called the credible intervals or
Bayesian confidence intervals.
In order to find the posterior probability density, we have to find the normalization factor (global likelihood), which is frequently difficult to calculate. In this
case the ratio of integrals of the likelihood over parameter intervals (subvolumes
in the multidimensional case) is used to find the odds that one parameter interval
is more likely than the other, given the observations.
In the multidimensional case, if we are interested only in a subset of model parameters, confidence values can be found by integrating the posterior distribution
over the remaining parameters; this is called marginalization.
403
394
395
Alternative
REJECT
ACCEPT
Hypothesis
ACCEPT
REJECT
t1
Figure B.1 Critical regions and type I and type II errors. The rejection level t1 divides
the sample space into critical regions. The probability of making a type I error is given by
the area under the probability density curve in the rejection region of the hypothesis in
the lower panel; it is the significance of the test. The power of the test is determined by
the probability of making a type II error (area under the probability density curve in the
rejection region of the alternative test in the upper panel); the power of the test is 1 .
population of cluster galaxies) may become large, constituting a large percentage of the true cluster population. A recent method that allows to control the
ratio of the number of type I errors (erroneously rejected field galaxies) to the
total number of rejections (the cluster galaxies) is called the false discovery
rate (FDR) (Nichol et al. 2000; Miller et al. 2001).
The generalized 2 test (Meyer 1975) measures the goodness-of-fit of the observed distribution of the sample to a given proposed distribution function of the
population. If the probability distribution function depends on unknown parameters the problem is a composite hypothesis test. As above, we consider a single
unknown parameter .
The N observation data points, fxi gN
with fj
i=1 , are binned into n intervals,
Pn
the number of observations lying in the j th interval. Therefore j =1 fj = N .
Once the parameter has been estimated, for example, by the maximum likelihood
method, it is possible to calculate the probability of observing a point x in the
interval j for the proposed distribution
pj (b) =
f (x; b)dx:
405
2002 by CRC Press LLC
396
Npj (b) is, therefore, the expected frequency of points lying in the interval j .
The statistic
n
X
[fj Npj (b)]2
X2 =
(B.7)
Npj (b)
j =1
is, for large values of n, approximately distributed according to a 2 distribution
function with n 2 degrees of freedom (the degrees of freedom are, in general,
n k 1, where k is the number of estimated parameters). The hypothesis that
the distribution of the population is the proposed one is rejected with a level of
significance if X 2 > 2;n k 1 .
A special kind of maximum likelihood method for parameter estimation is the
2 minimization method (Meyer 1975; Press et al. 1992). This is very popular
in cosmology, particularly in the analysis of the CMB data. The starting point
here is a set of N points with the form (xi ; yi ; i ). i is the rms deviation of the
measurement yi (xi ) that is assumed to be normally distributed with mean value
f (xi ; ) and variance i2 . The 2 statistic is defined as
N
X
[yi f (xi ; )]2
:
2 =
(B.8)
i2
i
406
2002 by CRC Press LLC
References
Abell G.O. (1958) The distribution of rich clusters of galaxies. Astrophys. J. Suppl. Ser. 3,
211288.
Abell G.O., Corwin H.G., Jr. and Olowin R.P. (1989) A catalog of rich clusters of galaxies.
Astrophys. J. Suppl. Ser. 70, 1138.
Abramowicz M.A. and Stegun I.A. (1965) Handbook of Mathematical Functions. Dover
Publications, New York.
Adler R.J. (1981) The Geometry of Random Fields. John Wiley & Sons, New York.
Aikio J. and Mahonen P. (1998) A simple void-searching algorithm. Astrophys. J. 497,
534540.
Alimi J.M., Valls-Gabaud D. and Blanchard A. (1988) A cross-correlation analysis of luminosity segregation in the clustering of galaxies. Astron. Astrophys. 206, L11L14.
Amendola L. and Palladino E. (1999) The scale of homogeneity in the Las Campanas
redshift survey. Astrophys. J. Lett. 514, L1L4.
Argoul F., Arneodo A., Elezgaray J. and Grasseau G. (1989) Wavelet transform of fractal
aggregates. Phys. Lett. A 135, 327336.
Arneodo A., Grasseau G. and Holschneider M. (1988) Wavelet transform of multifractals.
Phys. Rev. Lett. 61, 22812284.
Avnir D., Biham O., Lidar D. and Malcai O. (1998) Is the geometry of nature fractal?
Science 279, 3940.
Babu G.J. and Feigelson E.D. (1996a) Spatial point processes in astronomy. J. Statis. Planning Inference 50, 311326.
Babu G.J. and Feigelson E.D. (1996b) Astrostatistics. Chapman & Hall, London.
Baddeley A. (1999) Spatial sampling and censoring. In Stochastic Geometry. Likelihood and Computation (eds. O.E. Barndorff-Nielsen, W.S. Kendall and M.N.M. van
Lieshout), Chapman & Hall, Boca Raton, pp. 3778.
Baddeley A.J., Kerscher M., Schladitz K. and Scott B. (2000) Estimating the J function
without edge correction. Stat. Neerlandica 54, 114.
Baddeley A.J., Moyeed R.A., Howard C.V. and Boyde A.S. (1993) Analysis of a threedimensional point pattern with replication. Appl. Statist. 42, 641668.
Bahcall N.A. and Soneira R.M. (1983) The spatial correlation function of rich clusters of
galaxies. Astrophys. J. 270, 2038.
Bahcall N.A. and West M.J. (1992) The cluster correlation function: Consistent results
from an automated survey. Astrophys. J. 392, 419423.
Baleisis A., Lahav O., Loan A.J. and Wall J.V. (1998) Searching for large-scale structure
in deep radio surveys. Mon. Not. R. Astr. Soc. 297, 545558.
Balian R. and Schaeffer R. (1989) Scale-invariant matter distribution in the universe. I.
Counts in cells. Astron. Astrophys. 220, 129.
407
2002 by CRC Press LLC
398
References
Bardeen J.M., Bond J.R., Kaiser N. and Szalay A.S. (1986) The statistics of peaks of
Gaussian random fields. Astrophys. J. 304, 1561.
Barlett M.S. (1964) The spectral analysis of two-dimensional point processes. Biometrika
51, 299311.
Barlow R. (1989) Statistics. A Guide to the Use of Statistical Methods in Physical Sciences.
The Manchester Physics Series, John Wiley & Sons, Chichester.
Barnsley M. (1988) Fractals Everywhere. Academic Press, London.
Barrow J.D. (1992) Some statistical problems in cosmology with discussion by S.P.
Bhavsar and F.L. Bookstein. In Statistical Challenges in Modern Astronomy I (eds. G.J.
Babu and E.D. Feigelson), Springer-Verlag, New York, pp. 2155.
Barrow J.D., Sonoda D.H. and Bhavsar S.P. (1984) A bootstrap resampling analysis of
galaxy clustering. Mon. Not. R. Astr. Soc. 210, 19P23P.
Barrow J.D., Sonoda D.H. and Bhavsar S.P. (1985) Minimal spanning trees, filaments and
galaxy clustering. Mon. Not. R. Astr. Soc. 216, 1735.
Bartelmann M. and Schneider P. (2001) Weak gravitational lensing. Phys. Rep. 340, 291
472.
Baugh C.M. (1996) The real space correlation function measured from the APM galaxy
survey. Mon. Not. R. Astr. Soc. 280, 267275.
Baugh C.M. and Efstathiou G. (1993) The three-dimensional power spectrum measured
from the APM galaxy survey I. Use of the angular correlation function. Mon. Not. R.
Astr. Soc. 265, 145156.
Baugh C.M. and Efstathiou G. (1994) The three-dimensional power spectrum measured
from the APM galaxy survey II. Use of the two-dimensional power spectrum. Mon.
Not. R. Astr. Soc. 267, 323332.
Beisbart C. and Kerscher M. (2000) Luminosity and morphology dependent clustering of
galaxies. Astrophys. J. 545, 625.
Bentez N. and Sanz J.L. (1999) Measuring
=b with weak lensing. Astrophys. J. Lett. 525,
L1L4.
Bennett C.L., Banday A., Gorski K.M., Hinshaw G., Jackson P., Keegstra P., Kogut A.,
Smoot G.F., Wilkinson D.T. and Wright E.L. (1996) 4-year Cobe DMR cosmic microwave background observations: Maps and basic results. Astrophys. J. Lett. 464, L1
L4.
Benson A.J., Baugh C.M., Cole S., Frenk C.S. and Lacey C.G. (2000) The dependence
of velocity and clustering statistics on galaxy properties. Mon. Not. R. Astr. Soc. 316,
107119.
Bernardeau F. and van de Weygaert R. (1996) A new method for accurate estimation of
velocity field statistics. Mon. Not. R. Astr. Soc. 279, 693711.
Bertschinger E. (1987) Path integral methods for primordial density perturbations Sampling of constrained Gaussian random fields. Astrophys. J. Lett. 323, L103L106.
Bertschinger E. (1992) Large-scale structures and motions: Linear theory and statistics.
In New Insights into the Universe, Lecture Notes in Physics 408 (eds. V.J. Martnez,
M. Portilla and D. Saez), Springer-Verlag, Berlin, pp. 65126.
Bertschinger E. (1998) Simulations of structure formation in the universe. Ann. Rev. Astron.
Astrophys. 36, 599654.
Bertschinger E. and Dekel A. (1989) Recovering the full velocity and density fields from
large-scale redshiftdistance samples. Astrophys. J. Lett. 336, L5L8.
408
2002 by CRC Press LLC
References
399
Bhavsar S.P. and Ling E.N. (1988) Are the filaments real? Astrophys. J. Lett. 331, L63
L68.
Binggeli B., Tarenghi M. and Sandage A. (1990) The abundance and morphological segregation of dwarf galaxies in the field. Astron. Astrophys. 228, 4260.
Binney J. and Tremaine S. (1987) Galactic Dynamics. Princeton University Press, Princeton.
Bistolas V. and Hoffmann Y. (1998) Non-linear constrained realizations of the large scale
structure. Astrophys. J. 492, 439491.
Blumenfeld R. and Mandelbrot B. (1997) Levy dusts, MittagLeffler statistics, mass fractal
lacunarity, and perceived dimension. Phys. Rev. E 56, 112118.
Bond J.R., Cole S., Efstathiou G. and Kaiser N. (1991) Excursion set mass functions for
hierarchical Gaussian fluctuations. Astrophys. J. 379, 440460.
Bond J.R. and Efstathiou G. (1984) Cosmic background radiation anisotropies in universes
dominated by nonbaryonic dark matter. Astrophys. J. Lett. 285, L45L48.
Bond J.R. and Myers S.T. (1996a) The peakpatch picture of cosmic catalogs. I. Algorithms. Astrophys. J. Suppl. 103, 139.
Bond J.R. and Myers S.T. (1996b) The peakpatch picture of cosmic catalogs. II. Validation. Astrophys. J. Suppl. 103, 4162.
Bonometto S.A., Iovino A., Guzzo L., Giovanelli R. and Haynes M. (1993) Correlation
functions from the Perseus-Pisces redshift survey. Astrophys. J. 419, 451458.
Bonometto S.A. and Sharp N.A. (1980) On the derivation of higher order correlation functions. Astron. Astrophys. 92, 222224.
Borgani S. (1993) The multifractal behaviour of hierarchical density distributions. Mon.
Not. R. Astr. Soc. 260, 537549.
Borgani S. (1995) Scaling in the universe. Phys. Rep. 193, 1152.
Borgani S. and Guzzo L. (2001) X-ray clusters of galaxies as tracers of structure in the
Universe. Nature 409, 3945.
Borgani S., Martnez V.J., Perez M.A. and Valdarnini R. (1994) Is there any scaling in the
cluster distribution? Astrophys. J. 435, 3748.
Bouchet F.R., Colombi S., Hivon E. and Juszkiewicz R. (1995) Perturbative Lagrangian
approach to gravitational instability. Astr. Astrophys. 296, 575608.
Bohringer H., Schuecker P., Guzzo L., Collins C.A., Voges W., Schindler S., Neumann
D.M., Cruddace R.G., De Grandi S., Chincarini G., Edge A.C., MacGillivray H.T. and
Shaver P. (2001) The ROSAT-ESO flux limited X-ray (REFLEX) galaxy cluster survey.
I. The construction of the cluster sample. Astron. Astrophys. 369, 826850.
Braun J. and Sambridge M. (1995) A numerical method for solving partial differential
equations on highly irregular evolving grids. Nature 376, 655660.
Bridle S.L., Hobson M.P., Lasenby A.N. and Saunders R. (1998) A maximum-entropy
method for reconstructing the projected mass distribution of gravitational lenses. Mon.
Not. R. Astr. Soc. 299, 895904.
Broadhurst T.J., Ellis R.S., Koo D.C. and Szalay A.S. (1990) Large-scale distribution of
galaxies at the galactic poles. Nature 343, 726728.
Buchert T. (1992) Lagrangian theory of gravitational instability of Friedman-Lemaitre cosmologies and the Zeldovich approximation. Mon. Not. R. Astr. Soc. 254, 729737.
Buchert T. and Martnez V.J. (1993) Two-point correlation function in pancake models and
the fair sample hypothesis. Astrophys. J. 411, 485500.
409
2002 by CRC Press LLC
400
References
Bugayevskiy L.M. and Snyder J.P. (1995) Map Projections: A Reference Manual. Taylor
& Francis, London.
Bunde A. and Havlin S. (eds.) (1994) Fractals in Science. Springer-Verlag, Berlin.
Bunn E.F. and White M. (1997) The four-year COBE normalization and large-scale structure. Astrophys. J. 480, 621.
Burrus C.S., Gopinath R.A. and Guo H. (1998) Introduction to Wavelets and Wavelet
Transforms. A Primer. Prentice-Hall, Upper Saddle River.
Burstein D. and Heiles C. (1982) Reddenings derived from H I and galaxy counts Accuracy and maps. Astronom. J. 87, 11651189.
Buryak O. and Doroshkevich A. (1996) Correlation function as a measure of the structure.
Astron. Astrophys. 306, 18.
Canavezes A., Springel V., Oliver S.J., Rowan-Robinson M., Keeble O., White S.D.M.,
Saunders W., Efstathiou G., Frenk C.S., McMahon R.G., Maddox S., Sutherland W. and
Tadros H. (1998) The topology of the IRAS Point Source Catalogue Redshift Survey.
Mon. Not. R. Astr. Soc. 297, 777793.
Cappi A., Benoist C., da Costa L.N. and Maurogordato S. (1998) Is the Universe a fractal?
Results from the Southern Sky Redshift Survey 2. Astron. Astrophys. 335, 779788.
Carroll B.W. and Ostlie D.A. (1996) Modern Astrophysics. Addison-Wesley, Reading.
Catelan P., Lucchin F., Matarrese S. and Porciani C. (1998) The bias field of dark matter
halos. Mon. Not. R. Astr. Soc. 297, 692712.
Cen R. (1998) Gaussian peaks and clusters of galaxies. Astrophys. J. 509, 494516.
Christensen L.L. (1996) Compound Redshift Catalogues and their Application to Redshift Distortions of the Two-Point Correlation Function. Masters thesis, University of
Copenhagen, Copenhagen.
Chui C.K. (1992) An Introduction to Wavelets. Academic Press, San Diego.
Coleman P.H. and Pietronero L. (1992) The fractal structure of the Universe. Phys. Rep.
213, 311391.
Coles P. (1992) Analysis of patterns in galaxy clustering with discussions by F. L. Bookstein and N.K. Bose. In Statistical Challenges in Modern Astronomy I (eds. G.J. Babu
and E.D. Feigelson), Springer-Verlag, New York, pp. 5781.
Coles P. (1993) Galaxy formation with local bias. Mon. Not. R. Astr. Soc. 262, 10651075.
Coles P. and Barrow J.D. (1987) Non-Gaussian statistics and the microwave background
radiation. Mon. Not. R. Astr. Soc. 228, 407426.
Coles P. and Chiang L.Y. (2000) Non-linearity and non-Gaussianity through phase information. http://arXiv.org/abs/astro-ph/0010521.
Coles P. and Jones B. (1991) A lognormal model for the cosmological mass distribution.
Mon. Not. R. Astr. Soc. 248, 113.
Coles P. and Lucchin F. (1995) Cosmology: The Formation and Evolution of Cosmic Structure. John Wiley & Sons, New York.
Coles P., Pearson R.C., Borgani S., Plionis M. and Moscardini L. (1998) The cluster distribution as a test of dark matter models. IV: Topology and geometry. Mon. Not. R. Astr.
Soc. 294, 245258.
Colombi S., Pogosyan D. and Souradeep T. (2000) Tree structure of the percolating Universe. Phys. Rev. Let. 26, 55155518.
Couchman H.M.P., Thomas P.A. and Pearce F.R. (1995) Hydra: An adaptive-mesh implementation of P3 M-SPH. Astrophys. J. 452, 797813.
410
2002 by CRC Press LLC
References
401
Cox D.R. and Isham V. (1980) Point Processes. Chapman & Hall, London.
Cramer H. and Leadbetter M.R. (1967) Stationary and Related Stochastic Processes. Sample Function Properties and Their Applications. John Wiley & Sons, New York.
Cressie N. (1991) Statistics for Spatial Data. John Wiley & Sons, New York.
Croft R.A.C., Dalton G.B., Efstathiou G., Sutherland W.J. and Maddox S.J. (1997) The
richness dependence of galaxy cluster correlations: Results from a redshift survey of
rich APM clusters. Mon. Not. R. Astr. Soc. 291, 305313.
Croft R.A.C. and Efstathiou G. (1994) The correlation function of rich clusters of galaxies
in CDM-like models. Mon. Not. R. Astr. Soc. 267, 390400.
Croft R.A.C. and Gaztanaga E. (1997) Reconstruction of cosmological density and velocity
fields in the Lagrangian Zeldovich approximation. Mon. Not. R. Astr. Soc. 285, 793
805.
da Costa L.N., Geller M.J., Pellegrini P.S., Latham D.W., Fairall A.P., Marzke R.O.,
Willmer C.N.A., Huchra J.P., Calderon J.H., Ramella M. and Kurtz M.J. (1994) A complete southern sky redshift survey. Astrophys. J. Lett. 424, L1L4.
da Costa L.N., Pellegrini P.S., Davis M., Meiksin A., Sargent W.L.W. and Tonry J.L. (1991)
Southern Sky Redshift Survey The catalog. Astrophys. J. Suppl. Ser. 75, 935964.
da Costa L.N., Willmer C.N.A., Pellegrini P.S., Chaves O.L., Rite C., Maia M.A.G., Geller
M.J., Latham D.W., Kurtz M.J., Huchra J.P., Ramella M., Fairall A.P., Smith C. and
Lpari S. (1998) The Southern Sky Redshift Survey. Astronom. J. 116, 17.
Daley D.J. and Vere-Jones D. (1988) An Introduction to the Theory of Point Processes.
Springer-Verlag, New York.
Dalton G.B., Maddox S.J., Sutherland W.J. and Efstathiou G. (1997) The APM Galaxy
Survey V. Catalogues of galaxy clusters. Mon. Not. R. Astr. Soc. 289, 263284.
Davis M. (1997) Is the universe homogeneous on large scales? In Critical Dialogues in
Cosmology (ed. N. Turok), World Scientific, Singapore, pp. 1323.
Davis M. and Geller M.J. (1976) Galaxy correlations as a function of morphological type.
Astrophys. J. 208, 1319.
Davis M. and Huchra J.P. (1982) A survey of galaxy redshifts. III. The density field and
the induced gravity field. Astrophys. J. 254, 437450.
Davis M., Meiksin A., Strauss M.A., da Costa L.N. and Yahil A. (1988) On the universality
of the two-point galaxy correlation function. Astrophys. J. Lett. 333, L9L12.
Davis M. and Peebles P.J.E. (1983) A survey of galaxy redshifts V. The two-point position and velocity correlations. Astrophys. J. 267, 465482.
de Bernardis P., Ade P.A.R., Bock J.J., Bond J.R., Borrill J., Boscaleri A., Coble K., Contaldi C.R., Crill B.P., Troia G.D., Farese P., Ganga K., Giacometti M., Hivon E., Hristov
V.V., Iacoangeli A., Jaffe A.H., Jones W.C., Lange A.E., Martinis L., Masi S., Mason
P., Mauskopf P.D., Melchiorri A., Montroy T., Netterfield C.B., Pascale E., Piacentini
F., Pogosyan D., Polenta G., Pongetti F., Prunet S., Romeo G., Ruhl J.E. and Scaramuzzi F. (2001) Multiple peaks in the angular power spectrum of the cosmic microwave
background: Significance and consequences for cosmology. http://arXiv.org/abs/astroph/0105296.
de Lapparent V., Geller M.J. and Huchra J.P. (1986) A slice of the universe. Astrophys. J.
Lett. 302, L1L5.
de Lapparent V., Geller M.J. and Huchra J.P. (1989) The luminosity function for the CfA
redshift survey slices. Astrophys. J. 343, 117.
411
2002 by CRC Press LLC
402
References
de Vaucouleurs G. (1970) The case for a hierarchical cosmology. Science 167, 12031213.
Dekel A. (1997) Cosmological implications of large-scale flows. In Galaxy Scaling Relations: Origins, Evolution and Applications (eds. L. da Costa and A. Renzini), SpringerVerlag, Berlin, pp. 245285.
Dekel A. and Bertschinger E. (1990) Potential, velocity and density fields from sparse and
noisy redshiftdistance samples: Method. Astrophys. J. 364, 349369.
Dekel A., Blumenthal G.R., Primack J.R. and Stanhill D. (1992) Large-scale periodicity
and Gaussian fluctuations. Mon. Not. R. Astr. Soc. 257, 715730.
Dekel A., Eldar A., Kolatt T., Yahil A., Willick J.A., Faber S.M., Courteau S. and Burstein
D. (1999) POTENT reconstruction from Mark III velocities. Astrophys. J. 522, 138.
Dekel A. and Lahav O. (1999) Stochastic nonlinear galaxy biasing. Astrophys. J. 520, 24
34.
Dicke R.H., Peebles P.J.E., Roll P.G. and Wilkinson, D.T. (1965) Cosmic black-body radiation. Astrophys. J. 142, 414419.
Diggle P. (1983) Statistical Analysis of Point Processes. Chapman & Hall, London.
Diggle P. (1985) A kernel method for smoothing point process data. Appl. Statist. 34, 138
147.
Djorgovski S. and Davis M. (1987) Fundamental properties of elliptical galaxies. Astrophys. J. 313, 5968.
Dodelson S. and Gaztanaga E. (2000) Inverting the angular correlation function. Mon. Not.
R. Astr. Soc. 312, 774780.
Doguwa S. and Upton G.J.G. (1989) Edge-corrected estimators for the reduced second
moment measure of point processes. Biom. J. 31, 563576.
Domnguez-Tenreiro R., Gomez-Flechoso M.A. and Martnez V.J. (1994) Scaling analysis
of the distribution of galaxies in the CfA catalog. Astrophys. J. 424, 4258.
Domnguez-Tenreiro R. and Martnez V.J. (1989) Multidimensional analysis of the largescale segregation of luminosity. Astrophys. J. Lett. 339, L9L11.
Domnguez-Tenreiro R., Roy L.J. and Martnez V.J. (1992) On the multifractal character
of the Lorenz attractor. Prog. Theor. Phys. 87, 11071118.
Doroshkevich A.G. (1970) Spatial structure of perturbations and the origin of rotation of
galaxies in the fluctuation theory. Astrofizika 6, 581600.
Doroshkevich A.G., Muller V., Retzlaff J. and Turchaninov V. (1999) Superlarge-scale
structure in N-body simulations. Mon. Not. R. Astr. Soc. 306, 575591.
Durrer R., Eckmann J.P., Sylos Labini F., Montuori M. and Pietronero L. (1997) Angular
projections of fractal sets. Europhys. Lett. 40, 491496.
Efron B. and Tibshirani R. (1993) An Introduction to the Bootstrap. Chapman & Hall,
London.
Efstathiou G., Davis M., Frenk C.S. and White S.D.M. (1985) Numerical techniques for
large cosmological N-body simulations. Astrophys. J. Suppl. Ser. 57, 241260.
Efstathiou G., Ellis R.S. and Peterson B.A. (1988) Analysis of a complete galaxy redshift
survey. II. The field-galaxy luminosity function. Mon. Not. R. Astr. Soc. 232, 431
461.
Efstathiou G., Fall S.M. and Hogan C. (1979) Self-similar gravitational clustering. Mon.
Not. R. Astr. Soc. 189, 203220.
412
2002 by CRC Press LLC
References
403
Efstathiou G. and Moody S.J. (2001) Maximum likelihood estimates of the two- and threedimensional power spectra of the APM galaxy survey. Mon. Not. R. Astr. Soc. 325,
16031615.
Ehlers J., Geren P. and Sachs R.K. (1968) Isotropic solutions of the EinsteinLiouville
equations. J. Math. Phys. 9, 13441349.
Einasto J., Einasto M., Frisch P., Gottloeber S., Mueller V., Saar V., Starobinsky A.A.,
Tago E., Tucker D. and Andernach H. (1997a) The superclustervoid network II. An
oscillating cluster correlation function. Mon. Not. R. Astr. Soc. 289, 801812.
Einasto J., Einasto M., Gottloeber S., Mueller V., Saar V., Starobinsky A.A., Tago E.,
Tucker D., Andernach H. and Frisch P. (1997b) A 120 Mpc periodicity in the threedimensional distribution of galaxy superclusters. Nature 385, 139141.
Einasto J., Einasto M. and Gramann M. (1989) Structure and formation of superclusters.
IX Self-similarity of voids. Mon. Not. R. Astr. Soc. 238, 155177.
Einasto J. and Gramann M. (1993) Transition scale to a homogeneous universe. Astrophys.
J. 407, 443447.
Einasto J., Joeveer M. and Saar E. (1980) Structure of superclusters and supercluster formation. Mon. Not. R. Astr. Soc. 193, 353375.
Einasto J., Klypin A.A., Saar E. and Shandarin S.F. (1984) Structure of superclusters and
supercluster formation III. Quantitative study of the Local Supercluster. Mon. Not.
R. Astr. Soc. 206, 529558.
Einasto M., Tago E., Jaaniste J., Einasto J. and Andernach H. (1997) The superclustervoid network I. The supercluster catalogue and large-scale distribution. Astr. Astrophys.
Suppl. Ser. 123, 119133.
Eisenstein D.J. and Hu W. (1999) Power spectra for cold dark matter and its variants.
Astrophys. J. 511, 515.
El-Ad H. and Piran T. (1997) Voids in the large-scale structure. Astrophys. J. 491, 421435.
El-Ad H., Piran T. and da Costa L.N. (1997) A catalogue of the voids in the IRAS 1.2-Jy
survey. Mon. Not. R. Astr. Soc. 287, 790798.
Escalera E. and Mazure A. (1992) Wavelet analysis of subclustering An illustration,
Abell 754. Astrophys. J. 388, 2332.
Faber S.M. and Jackson R.E. (1976) Velocity dispersions and mass-to-light ratios for elliptical galaxies. Astrophys. J. 204, 668683.
Fairall A. (1998) Large-Scale Structures in the Universe. John Wiley & Sons in association
with Praxis Publishing, Chichester.
Falco E.E., Kurtz M.J., Geller M.J., Huchra J.P., Peters J., Berlind P., Mink D.J., Tokarz
S.P. and Elwell B. (1999) The Updated Zwicky Catalog (UZC). PASP 111, 438452.
Falconer K.J. (1985) The Geometry of Fractal Sets. Cambridge University Press, Cambridge.
Falconer K.J. (1990) Fractal Geometry. Mathematical Foundations and Applications. John
Wiley & Sons, Chichester.
Farge M. (1992) Wavelet transforms and their application to turbulence. Annu. Rev. Fluid
Mech. 24, 395457.
Feder J. (1988) Fractals. Plenum Press, New York.
Feldman H.A., Frieman J.A., Fry J.N. and Scoccimarro R. (2001) Constraints on galaxy
bias, matter density, and primordial non-Gaussianity from the PSCz galaxy redshift survey. Phys. Rev. Let. 86, 14341437.
413
2002 by CRC Press LLC
404
References
Feldman H.A., Kaiser N. and Peacock J.A. (1994) Power-spectrum analysis of threedimensional redshift surveys. Astrophys. J. 426, 2337.
Felten J.E. (1977) Study of the luminosity function for field galaxies. Astronom. J. 82,
861878.
Fiksel T. (1988) Edge-corrected density estimators for point processes. Statistics 19, 67
76.
Fisher K.B., Davis M., Strauss M.A., Yahil A. and Huchra J.P. (1993) The power spectrum
of IRAS galaxies. Astrophys. J. 402, 4257.
Fisher K.B., Huchra J.P., Strauss M.A., Davis M., Yahil A. and Schlegel D. (1995) The
IRAS 1.2 Jy survey: Redshift data. Astrophys. J. Suppl. Ser. 100, 69103.
Fisher K.B., Lahav O., Hoffman Y., Lynden-Bell D. and Zaroubi S. (1994) Wiener reconstruction of density, velocity, and potential fields from all-sky galaxy redshift surveys.
Mon. Not. R. Astr. Soc. 272, 885910.
Fisher K.B., Strauss M.A., Davis M., Yahil A. and Huchra J.P. (1992) The density evolution
of IRAS galaxies. Astrophys. J. 389, 188195.
Fixsen D.J., Cheng E.S., Gales J.M., Mather J.C., Shafer R.A. and Wright E.L. (1996) The
cosmic microwave background spectrum from the Full COBE FIRAS data set. Astrophys. J. 473, 576587.
Folkes S., Ronen S., Price I., Lahav O., Colless M., Maddox S., Deeley K., Glazebrook
K., Bland-Hawthorn J., Cannon R., Cole S., Collins C., Couch W., Driver S.P., Dalton
G., Efstathiou G., Ellis R.S., Frenk C.S., Kaiser N., Lewis I., Lumsden S., Peacock J.,
Peterson B.A., Sutherland W. and Taylor K. (1999) The 2dF Galaxy Redshift Survey:
spectral types and luminosity functions. Mon. Not. R. Astr. Soc. 308, 459472.
Freysz E., Pouligny B., Argoul F. and Arneodo A. (1990) Optical wavelet transform of
fractal aggregates. Phys. Rev. Lett. 64, 745748.
Fry J.N. (1984) The galaxy correlation function hierarchy in perturbation theory. Astrophys. J. 279, 499510.
Fry J.N. (1985) Cosmological density fluctuations and large-scale structure. From N-point
correlation functions to the probability distribution. Astrophys. J. 289, 1017.
Gamow G. (1948) The origin of elements and the separation of galaxies. Phys. Rev. 74,
505506.
Gaztanaga E. (1992) N-point correlation functions in the CfA and SSRS redshift distribution of galaxies. Astrophys. J. Lett. 398, L17L20.
Gaztanaga E. (1994) High-order galaxy correlation functions in the APM galaxy survey.
Mon. Not. R. Astr. Soc. 268, 913924.
Geller M.J. and Huchra J.P. (1989) Mapping the universe. Science 246, 897903.
Ghez J.M. and Vaienti S. (1989) On the wavelet analysis of multifractal sets. J. Stat. Phys.
57, 415420.
Giavalisco M., Steidel C.S., Adelberger K.L., Dickinson M.E., Pettini M. and Kellogg M.
(1998) The angular clustering of Lyman-break galaxies at redshift z 3. Astrophys. J.
503, 543552.
Giovanelli R. and Haynes M.P. (1991) Redshift surveys of galaxies. Ann. Rev. Astron.
Astrophys. 29, 499541.
Giovanelli R., Haynes M.P. and Chincarini G.L. (1986) Morphological segregation in the
Pisces-Perseus supercluster. Astrophys. J. 300, 7792.
414
2002 by CRC Press LLC
References
405
Gorski K. (1988) On the pattern of perturbations of the Hubble flow. Astrophys. J. Lett.
332, L7L11.
Gott J. R. I., Dickinson M. and Melott A.L. (1986) The sponge-like topology of large-scale
structure in the universe. Astrophys. J. 306, 341357.
Gramann M., Cen R. and Gott J.R. (1994) Recovering the real density field of galaxies
from redshift space. Astrophys. J. 425, 382391.
Groth E.J. and Peebles P.J.E. (1977) Statistical analysis of catalogs of extragalactic objects VII. Two- and three-point correlation functions for the high-resolution ShaneWirtanen catalog of galaxies. Astrophys. J. 217, 385405.
Gunn J.E. (1995) The Sloan Digital Sky Survey. In American Astronomical Society Meeting, vol. 186, p. 4405.
Guzzo L. (1997) Is the universe homogeneous? (On large scales). New Astronomy 2, 517
532.
Guzzo L., Bartlett J.G., Cappi A., Maurogordato S., Zucca E., Zamorani G., Balkowski C.,
Blanchard A., Cayatte V., Chincarini G., Collins C.A., Maccagni D., MacGillivray H.,
Merighi R., Mignoli M., Proust D., Ramella M., Scaramella R., Stirpe G.M. and Vettolani G. (2000) The ESO Slice Project (ESP) galaxy redshift survey. VII. The redshift
and real-space correlation functions. Astron. Astrophys. 355, 116.
Guzzo L., Iovino A., Chincarini G., Giovanelli R. and Haynes M.P. (1991) Scale-invariant
clustering in the large-scale distribution of galaxies. Astrophys. J. Lett. 382, L5L9.
Guzzo L., Strauss M.A., Fisher K.B., Giovanelli R. and Haynes M.P. (1997) Redshiftspace distortions and the real-space clustering of different galaxy types. Astrophys. J.
489, 3748.
Hadwiger H. (1957) Vorlesungen u ber Inhalt, Oberflache und Isoperimetrie. SpringerVerlag, Berlin.
Hall P. (1988) Introduction to the Theory of Coverage Processes. John Wiley & Sons,
Chichester.
Halsey T.C., Jensen M.H., Kadanoff L.P., Procaccia I. and Shraiman B.I. (1986) Fractal
measures and their singularities: The characterization of strange sets. Phys. Rev. A 33,
11411151.
Hamilton A.J.S. (1988) Evidence for biasing in the CfA survey. Astrophys. J. Lett. 331,
L59L62.
Hamilton A.J.S. (1992) Measuring Omega and the real correlation function from the redshift correlation function. Astrophys. J. Lett. 385, L5L8.
Hamilton A.J.S. (1993a)
from the anisotropy of the redshift correlation function in the
IRAS 2 Jansky survey. Astrophys. J. Lett. 406, L47L50.
Hamilton A.J.S. (1993b) Toward better ways to measure the galaxy correlation function.
Astrophys. J. 417, 1935.
Hamilton A.J.S. (1998) Linear redshift distortions: A review. In The Evolving Universe,
vol. 231 of Astrophysics and Space Science Library. Kluwer Academic Publishers, Dordrecht, pp. 185275.
Hamilton A.J.S. (2000) Uncorrelated modes of the nonlinear power spectrum. Mon. Not.
R. Astr. Soc. 312, 257284.
Hamilton A.J.S. and Culhane M. (1996) Spherical redshift distortions. Mon. Not. R. Astr.
Soc. 278, 7386.
415
2002 by CRC Press LLC
406
References
Hamilton A.J.S., Kumar P., Lu E. and Matthews A. (1991) Reconstructing the primordial spectrum of fluctuations of the universe from the observed nonlinear clustering of
galaxies. Astrophys. J. Lett. 374, L1L4.
Hamilton A.J.S. and Tegmark M. (2000) Decorrelating the power spectrum of galaxies.
Mon. Not. R. Astr. Soc. 312, 285294.
Hamilton A.J.S. and Tegmark M. (2000) The real space power spectrum of the PSCz survey
from 0.01 to 300 h Mpc 1 . http://arXiv.org/abs/astro-ph/0008392.
Hamilton A.J.S., Tegmark M. and Padmanabhan N. (2000) Linear redshift distortions and
power in the PSCz survey. Mon. Not. R. Astr. Soc. 317, L23L27.
Hanisch K.H. (1981) On classes of random sets and point processes. Serdica 7, 160167.
Harrison E. (2000) Cosmology. The Science of the Universe. Cambridge University Press,
Cambridge.
Hatton S. (1999) Approaching a homogeneous galaxy distribution: Results from the
StromloAPM redshift survey. Mon. Not. R. Astr. Soc. 310, 11281136.
Hauser M.G. and Peebles P.J.E. (1973) Statistical analysis of catalogs of extragalactic objects. II. The Abell catalog of rich clusters. Astrophys. J. 185, 757785.
Heavens A.F. and Taylor A.N. (1995) A spherical harmonic analysis of redshift space.
Mon. Not. R. Astr. Soc. 275, 483497.
Heck A. and Perdang J.M. (eds.) (1991) Applying Fractals in Astronomy, vol. 3 of Lecture
Notes in Physics. Springer-Verlag, Berlin.
Heinrich L., Korner R., Mehlhorn N. and Muche L. (1998) Numerical and analytical computation of some second-order characteristics of spatial Poisson-Voronoi tesselations.
Statistics 31, 235262.
Hentschel H.G.E. and Procaccia I. (1983) The infinite number of generalized dimensions
of fractals and strange attractors. Physica D 8, 435444.
Hermit S., Santiago B.X., Lahav O., Strauss M.A., Davis M., Dressler A. and Huchra J.P.
(1996) The two-point correlation function and morphological segregation in the Optical
Redshift Survey. Mon. Not. R. Astr. Soc. 283, 709720.
Hernquist L. (1987) Performance characteristics of tree codes. Astrophys. J. Suppl. Ser. 64,
715734.
Hockney R.W. and Eastwood J.W. (1988) Computer Simulation Using Particles. Adam
Hilger, Bristol.
Hoffman Y. and Ribak E. (1991) Constrained realizations of Gaussian random fields: A
simple algorithm. Astrophys. J. Lett. 380, L5L8.
Holschneider M. (1988) On the wavelet transformation of fractal objects. J. Stat. Phys. 50,
963993.
Hubble E. (1934) The distribution of extra-galactic nebulae. Astrophys. J. 79, 876.
Hubble E. and Humason M.L. (1931) The velocity-distance relation among extra-galactic
nebulae. Astrophys. J. 74, 4380.
Hubble E.P. (1936) The Realm of Nebulae. Yale University Press, New Haven.
Huchra J., Davis M., Latham D. and Tonry J. (1983) A survey of galaxy redshifts IV.
The data. Astrophys. J. Suppl. Ser. 52, 89119.
Huchra J.P., Geller M.J., de Lapparent V. and Corwin H. G. J. (1990) The CfA redshift
survey Data for the NGP + 30 zone. Astrophys. J. Suppl. Ser. 72, 433470.
Icke V. (1984) Voids and filaments. Mon. Not. R. Astr. Soc. 206, 1P3P.
416
2002 by CRC Press LLC
References
407
Icke V. and van de Weygaert R. (1987) Fragmenting the universe. Astron. Astrophys. 184,
1632.
Jenkins A., Frenk C.S., Pearce F.R., Thomas P.A., Colberg J.M., White S.D.M., Couchman
H.M.P., Peacock J.A., Efstathiou G. and Nelson A.H. (1998) Evolution of structure in
cold dark matter universes. Astrophys. J. 499, 2040.
Jensen L.G. and Szalay A.S. (1986) N -point correlations for biased galaxy formation.
Astrophys. J. Lett. 305, L5L9.
Jensen M.H., Paladin G. and Vulpiani A. (1991) Multiscaling in multifractals. Phys. Rev.
Lett. 67, 208211.
Jing Y.P. and Borner G. (1998) The three-point correlation function of galaxies determined
from the Las Campanas redshift survey. Astrophys. J. 503, 3747.
Jones B.J.T., Coles P. and Martnez V.J. (1992) Heterotopic clustering. Mon. Not. R. Astr.
Soc. 259, 146154.
Joyce M., Montuori M., Sylos Labini F. and Pietronero L. (1999) Comment on the paper
The ESO Slice Project [ESP] galaxy redshift survey: V. Evidence for a D=3 sample
dimensionality. Astr. Astrophys. 344, 387392.
Kaiser N. (1984) On the spatial correlations of Abell clusters. Astrophys. J. Lett. 284, L9
L12.
Kaiser N. (1986) A sparse-sampling strategy for the estimation of large-scale clustering
from redshift surveys. Mon. Not, R. Astr. Soc. 219, 785790.
Kaiser N. (1987) Clustering in real space and in redshift space. Mon. Not, R. Astr. Soc. 227,
121.
Kaiser N. and Peacock J.A. (1991) Power-spectrum analysis of one-dimensional redshift
surveys. Astrophys. J. 379, 482506.
Kaiser N. and Squires G. (1993) Mapping the dark matter with weak gravitational lensing.
Astrophys. J. 404, 441450.
Kauffmann G., Colberg J.M., Diaferio A. and White S.D.M. (1999) Clustering of galaxies
in a hierachical universe: I. Methods and results at z = 0. Mon. Not. R. Astr. Soc. 303,
188206.
Kendall M.G. and Stuart A. (1963) The Advanced Theory of Statistics. Distribution Theory,
vol. I. Charles Griffin & Co. Ltd., London.
Kendall M.G. and Stuart A. (1967) The Advanced Theory of Statistics. Inference and Relationship, vol. II. Charles Griffin & Co. Ltd., London.
Kepner J., Fan X., Bahcall N., Gunn J., Lupton R. and Xu G. (1999) An automated cluster
finder: The adaptive matched filter. Astrophys. J. 517, 7891.
Kerscher M. (1999) The geometry of second-order statistics Biases in common estimators. Astron. Astrophys. 343, 333347.
Kerscher M. (2000) Statistical analysis of large-scale structure in the Universe. In Statistical Physics and Spatial Statistics: The Art of Analyzing and Modelling Spatial Structures and Pattern Formation (eds. R.K. Mecke and D. Stoyan), Lecture Notes in Physics,
vol. 554, Springer-Verlag, Berlin, pp. 3671.
Kerscher M. (2001) Constructing, characterizing, and simulating Gaussian and higherorder point distributions. Phys. Rev. E 64, 056109.
Kerscher M. and Martnez V.J. (1998) Galaxy clustering and spatial point processes. Bull.
Int. Statist. Inst. 57-2, 363366.
417
2002 by CRC Press LLC
408
References
Kerscher M., Pons-Bordera M.J., Schmalzing J., Trasarti-Battistoni R., Buchert T.,
Martnez V.J. and Valdarnini R. (1999) A Global Descriptor of Spatial Pattern Interaction in the Galaxy Distribution. Astrophys. J. 513, 543548.
Kerscher M., Schmalzing J., Buchert T. and Wagner H. (1998) Fluctuations in the IRAS
1.2 Jy catalogue. Astron. Astrophys. 333, 112.
Kerscher M., Schmalzing J., Retzlaff J., Borgani S., Buchert T., Gottlober S., Muller V.,
Plionis M. and Wagner H. (1997) Minkowski functionals of Abell/ACO clusters. Mon.
Not. R. Astr. Soc. 284, 7384.
Kerscher M., Szapudi I. and Szalay A.S. (2000) A comparison of estimators for the twopoint correlation function. Astrophys. J. Lett. 535, L13L16.
Khintchine A.Y. (1955) Mathematical Methods of Queueing Theory (in Russian). Papers
of the Mathematical Institute of V.A. Steklov, Moscow.
King C.R. and Ellis R.S. (1985) The evolution of spiral galaxies and uncertainties in interpreting galaxy counts. Astrophys. J. 288, 456464.
Kirshner R.P. (1996) Galaxy redshift surveys. In Dark Matter in the Universe (eds.
S. Bonometto, J.R. Primack and A. Provenzale), IOS Press, Amsterdam, pp. 3648.
Kirshner R.P., Oemler A. J., Schechter P.L. and Shectman S.A. (1981) A million cubic
megaparsec void in Bootes. Astrophys. J. Lett. 248, L57L60.
Kirshner R.P., Oemler A. J., Schechter P.L. and Shectman S.A. (1987) A survey of the
Bootes void. Astrophys. J. 314, 493506.
Klypin A. and Shandarin S.F. (1993) Percolation technique for galaxy clustering. Astrophys. J. 413, 4858.
Kneib J.P., Ellis R.S., Smail I., Couch W.J. and Sharples R.M. (1996) Hubble space telescope observations of the lensing cluster Abell 2218. Astrophys. J. 471, 643656.
Kolatt T., Dekel A., Ganon G. and Willick J.A. (1996) Simulating our cosmological neighborhood: Mock catalogs for velocity analysis. Astrophys. J. 458, 419457.
Kruskal J.B. (1956) On the shortest spanning subtree of a graph and the traveling salesman
problem. Proc. Amer. Math. Soc. 7, 4850.
Krzewina L.G. and Saslaw W.C. (1996) Minimal spanning tree statistics for the analysis
of large-scale structure. Mon. Not. R. Astr. Soc. 278, 869876.
Sylos Labini F. and Montuori M. (1998) Scale invariant properties of the APMStromlo
survey. Astr. Astrophys. 331, 809814.
Lahav O., Lilje P.B., Primack J.R. and Rees M.J. (1991) Dynamical effects of the cosmological constant. Mon. Not. R. Astr. Soc. 251, 128136.
Landy S.D. and Szalay A.S. (1993) Bias and variance of angular correlation functions.
Astrophys. J. 412, 6471.
Layzer D. (1956) A new model for the distribution of galaxies in space. Astronom. J. 61,
383385.
Le Cam L. (1990) Maximum likelihood: an introduction. Int. Statist. Rev. 58, 153171.
Lehmann E.L. (1959) Testing Statistical Hypotheses. John Wiley & Sons, New York.
Lehmann E.L. and Casella G. (1998) Theory of Point Estimation. Springer-Verlag, New
York.
Lemson G. and Sanders R.H. (1991) On the use of the conditional density as a description
of galaxy clustering. Mon. Not. R. Astr. Soc. 252, 319328.
418
2002 by CRC Press LLC
References
409
Lin H., Kirshner R.P., Shectman S.A., Landy S.D., Oemler A., Tucker D.L. and Schechter
P.L. (1996) The luminosity function of galaxies in the Las Campanas redshift survey.
Astrophys. J. 464, 6078.
Loredo T.J. (1990) From Laplace to supernova SN 1987A: Bayesian inference in astrophysics. In Maximum Entropy and Bayesian Methods (ed. P.F. Foug`ere), Kluwer Academic Publishers, Dordrecht, pp. 81142.
Loveday J. (1996) The APM bright galaxy catalogue. Mon. Not. R. Astr. Soc. 278, 1025
1048.
Loveday J., Maddox S.J., Efstathiou G. and Peterson B.A. (1995) The StromloAPM redshift survey. 2: Variation of galaxy clustering with morphology and luminosity. Astrophys. J. 442, 457468.
Loveday J., Peterson B.A., Maddox S.J. and Efstathiou G. (1996) The StromloAPM redshift survey. IV. The redshift catalog. Astrophys. J. Suppl. Ser. 107, 201214.
Lucy L.B. (1974) An iterative technique for the rectification of observed distributions.
Astronom. J. 79, 745754.
Luo S. and Vishniac E. (1995) Three-dimensional shape statistics: Methodology. Astrophys. J. Suppl. Ser. 96, 429460.
Luo S., Vishniac E.T. and Martel H. (1996) Three-dimensional shape statistics: Applications. Astrophys. J. 468, 6274.
Lynden-Bell D., Faber S.M., Burstein D., Davies R.L., Dressler A., Terlevich R.J. and Wegner G. (1988) Spectroscopy and photometry of elliptical galaxies. V. Galaxy streaming
toward the new supergalactic center. Astrophys. J. 326, 1949.
Maddox S.J., Efstathiou G. and Sutherland W.J. (1996) The APM Galaxy Survey III.
An analysis of systematic errors in the angular correlation function and cosmological
implications. Mon. Not. R. Astr. Soc. 283, 12271263.
Maddox S.J., Efstathiou G., Sutherland W.J. and Loveday J. (1990a) The APM galaxy
survey I. APM measurements and star-galaxy separation. Mon. Not. R. Astr. Soc.
243, 692712.
Maddox S.J., Efstathiou G., Sutherland W.J. and Loveday J. (1990b) Galaxy correlations
on large scales. Mon. Not. R. Astr. Soc. 242, 43P47P.
Mandelbrot B.B. (1975) Sur un mod`ele decomposable dUnivers hierarchise: deduction
des correlations galactiques sur la sph`ere celeste. C.R. Acad. Sci. (Paris) A 280, 1551
1554.
Mandelbrot B.B. (1982) The Fractal Geometry of Nature. Revised edition of Fractals
(1977). W.H. Freeman, San Francisco.
Mann R.G., Heavens A.F. and Peacock J.A. (1993) The richness dependence of cluster
correlations. Mon. Not. R. Astr. Soc. 263, 798816.
Margon B. (1999) The Sloan Digital Sky Survey. Phil. Trans. R. Soc. London A 357, 93
103.
Marriott F.H.C. (1990) A Dictionary of Statistical Terms. 5th ed., Longman Scientific &
Technical, Harlow.
Martel H. (1991) Linear perturbation theory and spherical overdensities in 6= 0 Friedmann models. Astrophys. J. 377, 713.
Martnez V.J. (1991) Fractal aspects of galaxy clustering. In Applying Fractals in Astronomy, vol. 3 of Lecture Notes in Physics (eds. A. Heck and J.M. Perdang J.M.), SpringerVerlag, Berlin, pp. 135159.
419
2002 by CRC Press LLC
410
References
Martnez V.J. (1996) Measures of galaxy clustering. In Dark Matter in the Universe (eds.
S. Bonometto, J.R. Primack and A. Provenzale), IOS Press, Amsterdam, pp. 255267.
Martnez V.J. (1997) Recent advances in large-scale structure statistics. In Statistical Challenges in Modern Astronomy II (eds. G.J. Babu and E.D. Feigelson), Springer-Verlag,
New York, pp. 153171.
Martnez V.J. (1999) Is the universe fractal? Science 284, 445446.
Martnez V.J. (Non)-fractality on large scales. In IAU Symp. 201: New Cosmological Data
and the Value of the Fundamental Parameters, ASP Conference Series. (eds. A. Lasenby
and A. Wilkinson), vol. 201, in press.
Martnez V.J. and Coles P. (1994) Correlations and scaling in the QDOT redshift survey.
Astrophys. J. 437, 550555.
Martnez V.J. and Jones B.J.T. (1990) Why the universe is not a fractal. Mon. Not. R. Astr.
Soc. 242, 517521.
Martnez V.J., Jones B.J.T., Domnguez-Tenreiro R. and van de Weygaert R. (1990) Clustering paradigms and multifractal measures. Astrophys. J. 357, 5061.
Martnez V.J., Lopez-Mart B. and Pons-Bordera M.J. (2001) Does the galaxy correlation
length increase with the sample depth? Astrophys. J. Lett. 554, L5L8.
Martnez V.J., Paredes S., Borgani S. and Coles P. (1995) Multiscaling properties of largescale structure in the universe. Science 269, 12451247.
Martnez V.J., Paredes S. and Saar E. (1993) Wavelet analysis of the multifractal character
of the galaxy distribution. Mon. Not. R. Astr. Soc. 260, 365375.
Martnez V.J., Pons-Bordera M., Moyeed R.A. and Graham M.J. (1998) Searching for the
scale of homogeneity. Mon. Not. R. Astr. Soc. 298, 12121222.
Martnez V.J., Portilla M., Jones B.J.T. and Paredes S. (1993) The galaxy clustering correlation length. Astron. Astrophys. 280, 519.
Matarrese S., Verde L. and Heavens A.F. (1997) Large-scale bias in the Universe: Bispectrum method. Mon. Not. R. Astr. Soc. 290, 651662.
Matern B. (1960) Spatial Variation. Medd. fran Statens Skogsforskningsinst. 49, 1144.
Matsubara T., Szalay A.S. and Landy S.D. (2000) Cosmological parameters from the
eigenmode analysis of the Las Campanas redshift survey. Astrophys. J. Lett. 535, L1L4.
Maurogordato S. and Lachi`eze-Rey M. (1987) Void probabilities in the galaxy distribution
Scaling and luminosity segregation. Astrophys. J. 320, 1325.
McCauley J.L. (1998) The galaxy distributions: Homogeneous, fractal, or neither? Fractals
6, 109119.
McCracken H.J., Metcalfe N., Shanks T., Campos A., Gardner J.P. and Fong R. (2000)
Galaxy number counts IV. Surveying the Herschel deep field in the near-infrared.
Mon. Not. R. Astr. Soc. 311, 707718.
Meakin P. (1987) Diffusion-limited aggregation on multifractal lattices: A model for fluidfluid displacement in porous media. Phys. Rev. A 36, 28332837.
Mecke K.R., Buchert T. and Wagner H. (1994) Robust morphological measures for largescale structure in the Universe. Astr. Astrophys. 288, 697704.
Mecke K.R. and Wagner H. (1991) Euler characteristic and related measures for random
geometric sets. J. Stat. Phys. 64, 843850.
Melott A.L. (1990) The topology of the large-scale structure in the universe. Phys. Rep.
193, 139.
420
2002 by CRC Press LLC
References
411
Melott A.L. and Dominik K.G. (1993) Tests of smoothing methods for topological study
of galaxy redshift surveys. Astrophys. J. Suppl. Ser. 86, 14.
Meyer S.L. (1975) Data Analysis for Scientists and Engineers. John Wiley & Sons, New
York.
Miller C.J. and Batuski D.J. (2001) The power spectrum of rich clusters on near-gigaparsec
scales. Astrophys. J. 551, 635642.
Miller C.J., Genovese C., Nichol R.C., Wasserman L., Connolly A., Reichart D., Hopkins
A., Schneider J. and Moore A. (2001) Controlling the false discovery rate in astrophysical data analysis. http://arXiv.org/abs/astro-ph/0107034.
Miller C.J., Nichol R.C. and Batuski D.J. (2001) Possible detection of baryonic fluctuations
in the large-scale structure power spectrum. Astrophys. J. 555, 6873.
Milne E.A. (1933) World-structure and the expansion of the universe. Zeitschr. fur Astrophys. 6, 135.
Mo H.J., Jing Y.P. and Borner G. (1992) On the error estimates of correlation functions.
Astrophys. J. 392, 452457.
Mo H.J., Peacock J.A. and Xia X.Y. (1993) The cross-correlation of IRAS galaxies with
Abell clusters and radio galaxies. Mon. Not. R. Astr. Soc. 260, 121131.
Mo H.J. and White S.D.M. (1996) An analytic model for the spatial clustering of dark
matter haloes. Mon. Not. R. Astr. Soc. 282, 347361.
Mo H.J., Ying Y.P. and White S.D.M. (1997) High-order correlations of peaks and halos:
A step toward understanding galaxy biasing. Mon. Not. R. Astr. Soc. 284, 189201.
Mller J., Syversveen A.R. and Waagepetersen R.P. (1998) Log Gaussian Cox processes.
Scand. J. Statist. 25, 451482.
Mollerach S., Martnez V.J., Diego J.M., Martnez-Gonzalez E., Sanz J.L. and Paredes S.
(1999) The roughness of the last scattering surface. Astrophys. J. 525, 1724.
Monaco P. (1999) Dynamics in the cosmological mass function (or, why does the Press &
Schechter work?). In Observational Cosmology: The Development of Galaxy Systems
(eds. G. Giuricin, M. Mezzetti and P. Salucci), vol. 176, Astronomical Society of the
Pacific, San Francisco, pp. 186196.
Monaco P. and Efstathiou G. (1999) Reconstruction of cosmological initial conditions from
galaxy redshift catalogues. Mon. Not. R. Astr. Soc. 308, 763779.
Monaco P., Efstathiou G., Maddox S.J., Branchini E., Frenk C.S., McMahon R.G., Oliver
S.J., Rowan-Robinson M., Saunders W., Sutherland W.J., Tadros H. and White S.D.M.
(2000) The 1-point PDF of the initial conditions of our local universe from the IRAS
PSC redshift catalogue. Mon. Not. R. Astr. Soc. 318, 681692.
Monaghan J.J. (1992) Smoothed particle hydrodynamics. Ann. Rev. Astron. Astrophys. 30,
543574.
Monin A.S. and Yaglom A.M. (1975) Statistical Fluid Mechanics. MIT Press, Cambridge.
Moore B., Frenk C.S., Weinberg D.H., Saunders W., Lawrence A., Ellis R.S., Kaiser N.,
Efstathiou G. and Rowan-Robinson M. (1992) The topology of the QDOT IRAS redshift
survey. Mon. Not. R. Astr. Soc. 256, 477499.
Muller V., Arbabi-Bidgoli S., Einasto J. and Tucker D. (2000) Voids in the Las Campanas
redshift survey versus cold dark matter models. Mon. Not. R. Astr. Soc. 318, 280288.
Narayan R. and Bartelmann M. (1999) Lectures on gravitational lensing. In Formation of
Structure in the Universe, Cambridge University Press, pp. 360421.
421
2002 by CRC Press LLC
412
References
Narayanan V.K. and Croft R.A.C. (1999) Recovering the primordial fluctuations: A comparison of methods. Astrophys. J. 515, 471486.
Neyman J. and Scott E.L. (1952) A theory of the spatial distribution of galaxies. Astrophys.
J. 116, 144163.
Neyman J. and Scott E.L. (1955) On the inapplicability of the theory of fluctuations to
galaxies. Astronom. J. 60, 33.
Neyman J. and Scott E.L. (1958) Statistical approach to problems of cosmology. J. Roy.
Statist. Soc. B 20, 143.
Nichol R., Miller C.J., Connolly A., Chong S.S., Genovese C., Moore A., Reichart D.,
Schneider J., Wasserman L., Annis J., Brinkman J., Bohringer H., Castander F., Kim R.,
McKay T., Postman M., Sheldon E., Szapudi I., Romer K. and Voges W. (2000) SDSS
RASS: Next generation of cluster-finding algorithms. In http://arXiv.org/abs/astro-ph/
0011557.
Nilson P. (1973) Uppsala general catalogue of galaxies. Nova Acta Regiae Soc. Sci. Upsaliensis Ser. V 1, 1447.
Nusser A. and Dekel A. (1990) Filamentary structure from Gaussian fluctuations using the
adhesion approximation. Astrophys. J. 362, 1424.
Nusser A., Dekel A., Bertschinger E. and Blumenthal G. (1991) Cosmological velocity
density relation in the quasi-linear regime. Astrophys. J. 379, 618.
Nusser A. and Lahav O. (2000) The Lyman- forest in a truncated hierarchical structure
formation. Mon. Not. R. Astr. Soc. 313, L39L42.
Ohser J. and Stoyan D. (1981) On the second-order and orientation analysis of planar
stationary point processes. Biom. J. 23, 523533.
Okabe A., Boots B., Sugihara K. and Chiu S.N. (1999) Spatial Tessellations. Concepts and
Applications of Voronoi Diagrams. John Wiley & Sons, Chichester.
Padmanabhan N., Tegmark M. and Hamilton A.J.S. (2000) The power spectrum of the
CfA/SSRS UZS galaxy redshift survey. Astrophys. J. 550, 5264.
Padmanabhan T. (1993) Structure Formation in the Universe. Cambridge University Press,
Cambridge.
Pan J. and Coles P. (2000) Large-scale cosmic homogeneity from a multifractal analysis of
the PSCz catalogue. Mon. Not. R. Astr. Soc. 318, L51L54.
Paredes S., Jones B.J.T. and Martnez V.J. (1995) The clustering of galaxy clusters
Synthetic distributions and the correlation function amplitude. Mon. Not. R. Astr. Soc.
276, 11161130.
Park C., Vogeley M.S., Geller M.J. and Huchra J.P. (1994) Power spectrum, correlation
function, and tests for luminosity bias in the CfA redshift survey. Astrophys. J. 431,
569585.
Paul T. (1985) Ph.D. thesis, Univ. Aix-Marseille II, Luminy, Marceille, France.
Peacock J.A. (1999) Cosmological Physics. Cambridge University Press, Cambridge.
Peacock J.A. (2000) Clustering of mass and galaxies. In http://arXiv.org/abs/astro-ph/
0002013.
Peacock J.A., Cole S., Norberg P., Baugh C.M., Bland-Hawthorn J., Bridges T., Cannon
R.D., Colless M., Collins C., Couch W., Dalton G., Deeley K., Propris R.D., Driver
S.P., Efstathiou G., Ellis R.S., Frenk C.S., Glazebrook K., Jackson C., Lahav O., Lewis
I., Lumsden S., Maddox S., Percival W.J., Peterson B.A., Price I., Sutherland W. and
Taylor K. (2001) A measurement of the cosmological mass density from clustering in
the 2dF galaxy redshift survey. Nature 410, 169173.
422
2002 by CRC Press LLC
References
413
Peacock J.A. and Dodds S.J. (1996) Non-linear evolution of cosmological power spectra.
Mon. Not. R. Astr. Soc. 280, L19L26.
Peacock J.A. and Heavens A.F. (1990) Alternatives to the PressSchechter cosmological
mass function. Mon. Not. R. Astr. Soc. 243, 133143.
Peacock J.A. and Smith R.E. (2000) Halo occupation numbers and galaxy bias. Mon. Not.
R. Astr. Soc. 318, 11441156.
Peebles P.J.E. (1973) Statistical analysis of catalogs of extragalactic objects. I. Theory.
Astrophys. J. 185, 413440.
Peebles P.J.E. (1978) Stability of a hierarchical clustering pattern in the distribution of
galaxies. Astron. Astrophys. 68, 345352.
Peebles P.J.E. (1980) The Large-Scale Structure of the Universe. Princeton University
Press, Princeton.
Peebles P.J.E. (1989) Tracing galaxy orbits back in time. Astrophys. J. Lett. 344, L53L56.
Peebles P.J.E. (1993) Principles of Physical Cosmology. Princeton University Press,
Princeton.
Peebles P.J.E. (1998) The standard cosmological model. In Les Rencontres de Physique
de la Vallee dAoste, Results and Perspectives in Particle Physics (ed. M. Greco),
Poligrafica Laziale s.r.l., Frascati, pp. 3956.
Peebles P.J.E. (2001) The galaxy and mass N-Point correlation functions: a blast from the
past. In Historical Development of Modern Cosmology, ASP Conference Series (eds.
V.J. Martnez, V. Trimble and M.J. Pons-Bordera), vol. 252, Astronomical Society of
the Pacific, San Francisco, pp. 201218.
Peebles P.J.E. and Hauser M.G. (1974) Statistical analysis of catalogs of extragalactic objects. III. The Shane-Wirtanen and Zwicky catalogs. Astrophys. J. Suppl. Ser. 28, 1936.
Peitgen H.O. and Saupe D. (eds.) (1988) The Science of Fractal Images. Springer-Verlag,
New York.
Penzias A.A. and Wilson, R.W. (1965) A measurement of excess antenna temperature at
4080 Mc/s. Astrophys. J. 142, 419420.
Percival W.J., Baugh C.M., Bland-Hawthorn J., Bridges T., Cannon R., Cole S., Colless
M., Collins C., Couch W., Dalton G., Propris R.D., Driver S.P., Efstathiou G., Ellis
R.S., Frenk C.S., Glazebrook K., Jackson C., Lahav O., Lewis I., Lumsden S., Maddox
S., Moody S., Norberg P., Peacock J.A., Peterson B.A., Sutherland W. and Taylor K.
(2001) The 2dF galaxy redshift survey: The power spectrum and the matter content of
the universe. http://arXiv.org/abs/astro-ph/0105252.
Perlmutter S., Aldering G., Goldhaber G., Knop R.A., Nugent P., Castro P.G., Deustua S.,
Fabbro S., Goobar A., Groom D.E., Hook I.M., Kim A.G., Kim M.Y., Lee J.C., Nunes
N.J., Pain R., Pennypacker C.R., Quimby R., Lidman C., Ellis R.S., Irwin M., McMahon
R.G., Ruiz-Lapuente P., Walton N., Schaefer B., Boyle B.J., Filippenko A.V., Matheson
T., Fruchter A.S., Panagia N., Newberg H.J.M. and Couch W.J. (1999) Measurements
of
and from 42 high-redshift supernovae. Astrophys. J. 516, 565586.
Phillipps S. and Shanks T. (1987) On the variation of galaxy correlations with luminosity.
Mon. Not. R. Astr. Soc. 229, 621626.
Pierpaoli E., Scott D. and White M. (2001) Power spectrum normalization from the local
abundance of rich clusters of galaxies. Mon. Not. R. Astr. Soc. 325, 7788.
Pietronero L. (1987) The fractal structure of the Universe: correlations of galaxies and
clusters and the average mass density. Physica A 144, 257284.
423
2002 by CRC Press LLC
414
References
424
2002 by CRC Press LLC
References
415
Rowan-Robinson M., Saunders W., Lawrence A. and Leech K. (1991) The QMW IRAS
galaxy catalogue A highly complete and reliable IRAS 60-micron galaxy catalogue.
Mon. Not. R. Astr. Soc. 253, 485495.
Rybicki G.B. and Press W.H. (1992) Interpolation, realization, and reconstruction of noisy,
irregularly sampled data. Astrophys. J. 398, 169176.
Saar E. (1973) Evolution of density contrast in underdense regions. Publ. Tartu Astrophys.
Obs. 41, 3040.
Sahni V. and Coles P. (1995) Approximation methods for non-linear gravitational clustering. Phys. Reports 262, 1136.
Sahni V., Sathyaprakash B.S. and Shandarin S.F. (1997) Probing large scale structure using
percolation and genus curves. Astrophys. J. Lett. 476, L1L6.
Sahni V., Sathyaprakash B.S. and Shandarin S.F. (1998) Shapefinders: a new shape diagnostic for large-scale structure. Astrophys. J. Lett. 495, L5L8.
Salzer J.J., Hanson M.M. and Gavazzi G. (1990) The relative spatial distributions of highand low-luminosity galaxies toward Coma. Astrophys. J. 353, 3950.
Santalo L. (1976) Integral Geometry and Geometric Probability. Addison-Wesley, Reading.
Sanz J.L., Herranz D. and Martnez-Gonzalez E. (2001) Optimal detection of sources on a
homogeneous and isotropic background. Astrophys. J. 552, 484492.
Saslaw W.C. (2000) The Distribution of the Galaxies. Gravitational Clustering in Cosmology. Cambridge University Press, Cambridge.
Saunders W. and Ballinger B.E. (2000) Interpolation of discretely-sampled density fields.
In http://arXiv.org/abs/astro-ph/0005606.
Saunders W., Frenk C., Rowan-Robinson M., Lawrence A. and Efstathiou G. (1991) The
density field of the local universe. Nature 349, 3238.
Saunders W., Rowan-Robinson M., Lawrence A., Efstathiou G., Kaiser N., Ellis R.S. and
Frenk C.S. (1990) The 60-m and far-infrared luminosity functions of the IRAS galaxies. Mon. Not. R. Astr. Soc. 242, 318337.
Saunders W., Sutherland W.J., Maddox S.J., Keeble O., Oliver S.J., Rowan-Robinson M.,
McMahon R.G., Efstathiou G.P., Tadros H., White S.D.M., Frenk C.S., Carraminana A.
and Hawkins M.R.S. (2000) The PSCz catalogue. Mon. Not. R. Astr. Soc. 317, 5564.
Scaramella R., Guzzo L., Zamorani G., Zucca E., Balkowski C., Blanchard A., Cappi A.,
Cayatte V., Chincarini G., Collins C., Fiorani A., Maccagni D., MacGillivray H., Maurogordato S., Merighi R., Mignoli M., Proust D., Ramella M., Stirpe G.M. and Vettolani
G. (1998) The ESO Slice Project [ESP] galaxy redshift survey: V. Evidence for a D=3
sample dimensionality. Astr. Astrophys. 334, 404408.
Schaap W.E. and van de Weygaert R. (2000) Continuous fields and discrete samples: reconstruction through Delaunay tessellations. Astron. Astrophys. 363, L29L32.
Schaap W.E. and van de Weygaert R. (2001) Delaunay recovery of cosmic density and
velocity probes. http://arXiv.org/abs/astro-ph/0109261.
Schechter P. (1976) An analytic expression for the luminosity function for galaxies. Astrophys. J. 203, 297306.
Scherrer R.J. and Bertschinger E. (1991) Statistics of primordial density perturbations from
discrete seed masses. Astrophys. J. 381, 349360.
425
2002 by CRC Press LLC
416
References
Schlegel D.J., Finkbeiner D.P. and Davis M. (1998) Maps of dust infrared emission for use
in estimation of reddening and cosmic microwave background radiation foregrounds.
Astrophys. J. 500, 525553.
Schmalzing J. and Gorski K.M. (1998) Minkowski functionals used in the morphological
analysis of cosmic microwave background anisotropy maps. Mon. Not. R. Astr. Soc. 297,
355365.
Schmalzing J., Kerscher M. and Buchert T. (1996) Minkowski functionals in cosmology.
In Dark Matter in the Universe (eds. S. Bonometto, J.R. Primack and A. Provenzale),
IOS Press, Amsterdam, pp. 281291.
Schmoldt I.M., Saar V., Saha P., Branchini E., Efstathiou G.P., Frenk C.S., Keeble O.,
Maddox S., McMahon R., Oliver S., Rowan-Robinson M., Saunders W., Sutherland
W.J., Tadros H. and White S.D.M. (1999) On density and velocity fields and from the
IRAS PSCz survey. Astronom. J. 118, 11461160.
Schneider P., Ehlers J. and Falco E.E. (1992) Gravitational Lenses. Springer-Verlag, New
York.
Schneider P., van Waerbeke L., Jain B. and Kruse G. (1998) A new measure for cosmic
shear. Mon. Not. R. Astr. Soc. 296, 873892.
Schramm D.N. and Turner M.S. (1998) Big-bang nucleosynthesis enters the precision era.
Rev. Mod. Phys. 70, 303318.
Scoccimarro R. (2000) The bispectrum: From theory to observations. Astrophys. J. 544,
597615.
Scoccimarro R., Feldman H.A., Fry J.N. and Frieman J.A. (2001) The bispectrum of IRAS
redshift catalogs. Astrophys. J. 546, 652664.
Scoccimarro R. and Frieman J.A. (1999) Hyperextended cosmological perturbation theory:
Predicting non-linear clustering amplitudes. Astrophys. J. 520, 3544.
Scoccimarro R. and Sheth R.K. (2001) PTHalos: A fast method for generating mock galaxy
distributions. http://arXiv.org/abs/astro-ph/0106120.
Seldner M., Siebers B., Groth E.J. and Peebles P.J.E. (1977) New reduction of the Lick
catalog of galaxies. Astronom. J. 82, 249256.
Seljak U. (1998) Cosmography and power spectrum estimation: A unified approach. Astrophys. J. 503, 492501.
Seljak U. (2000) Analytic model for galaxy and dark matter clustering. Mon. Not. R. Astr.
Soc. 318, 203213.
Seljak U. (2001) Redshift space bias and from the halo model. Mon. Not. R. Astr. Soc.
325, 13591364.
Seljak U. and Zaldarriaga M. (1996) A line-of-sight integration approach to cosmic microwave background anisotropies. Astrophys. J. 469, 437444.
Shandarin S.F. (2001) Testing non-Gaussianity in CMB maps by morphological statistics.
http://arXiv.org/abs/astro-ph/0107319.
Shandarin S.F. (1983) Percolation theory and the cell-lattice structure of the universe. Sov.
Astr. Lett. 9, 100102.
Shane C.D. and Wirtanen C.A. (1967) The Lick catalog of galaxies. Publ. Lick Obs. 22, 1.
Shanks T. (1990) Galaxy count models and the extragalactic background light. In IAU
Symp. 139: The Galactic and Extragalactic Background Radiation (eds. S. Bowyer and
C. Leinert), vol. 139, pp. 269281.
426
2002 by CRC Press LLC
References
417
Sheth R.K. (1995) Constrained realizations and minimum variance reconstruction of nonGaussian random fields. Mon. Not. R. Astr. Soc. 277, 933944.
Sheth R.K. and Lemson G. (1999) The forest of merger history trees associated with the
formation of dark matter haloes. Mon. Not. R. Astr. Soc. 305, 946956.
Sheth R.K., Mo H.J. and Tormen G. (2001) Ellipsoidal collapse and an improved model
for the number and spatial distribution of dark matter haloes. Mon. Not. R. Astr. Soc.
323, 112.
Sheth R.K. and Tormen G. (1996) Large scale bias and the peak background split. Mon.
Not. R. Astr. Soc. 308, 119126.
Sibson R. (1981) A brief description of natural neighbor interpolations. In Interpreting
Multivariate Data (ed. V. Barnett), John Wiley & Sons, Chichester, pp. 153171.
Sigad Y., Eldar A., Dekel A., Strauss M.A. and Yahil A. (1998) IRAS versus POTENT
density fields on large scales: Biasing and Omega. Astrophys. J. 495, 516532.
Silberman L., Dekel A., Eldar A. and Zehavi I. (2001) Cosmological density and power
spectrum from peculiar velocities: Nonlinear corrections and principal component analysis. Astrophys. J. 557, 102116.
Silverman B.W. (1986) Density Estimation for Statistics and Data Analysis. Chapman &
Hall, London.
Slezak E., Bijaoui A. and Mars G. (1990) Identification of structures from galaxy counts
Use of the wavelet transform. Astron. Astrophys. 227, 301316.
Slezak E., de Lapparent V. and Bijaoui A. (1993) Objective detection of voids and highdensity structures in the first CfA redshift survey slice. Astrophys. J. 409, 517529.
Snethlage M. (1999) Is bootstrap really helpful in point process statistics? Metrika 49,
245255.
Snethlage M., Martnez V.J., Stoyan D. and Saar E. Point field models for the galaxy point
pattern. Astron. Astrophys., submitted.
Soneira R.M. and Peebles P.J.E. (1978) A computer model universe Simulation of the
nature of the galaxy distribution in the Lick catalog. Astronom. J. 83, 845849.
Starck J.L., Murtagh F. and Bijaoui A. (1998) Image Processing and Data Analysis. The
Multiscale Approach. Cambridge University Press, Cambridge.
Stein M.L. (1993) Asymptocially optimal estimation of the reduced second moment measure of point processes. Biometrika 80, 443449.
Stein M.L. (1997) Discussion by Michael L. Stein. In Statistical Challenges in Modern
Astronomy II (eds. G.J. Babu and E.D. Feigelson), Springer-Verlag, New York, pp. 166
171.
Stoyan D. (1984a) On correlations of marked point processes. Math. Nachr. 116, 197207.
Stoyan D. (1984b) Correlations of the marks of marked point processes statistical inference and simple models. J. Inf. Process. Cybern. 20, 285294.
Stoyan D. (1994) Caution with fractal point patterns! Statistics 25, 267270.
Stoyan D. (2000) Basic ideas of spatial statistics. In Statistical Physics and Spatial Statistics: The Art of Analyzing and Modelling Spatial Structures and Pattern Formation,
Lecture Notes in Physics (eds. R.K. Mecke and D. Stoyan), vol. 554, Springer-Verlag,
Berlin, pp. 321.
Stoyan D., Kendall W.S. and Mecke J. (1995) Stochastic Geometry and Its Applications.
John Wiley & Sons, Chichester.
427
2002 by CRC Press LLC
418
References
Stoyan D. and Stoyan H. (1994) Fractals, Random Shapes and Point Fields. John Wiley &
Sons, Chichester.
Stoyan D. and Stoyan H. (2000) Improving ratio estimators of second-order point process
characteristics. Scand. J. Statist. 27, 641656.
Strauss M.A., Davis M., Yahil A. and Huchra J.P. (1990) A redshift survey of IRAS galaxies I. Sample selection. Astrophys. J. 361, 4962.
Strauss M.A., Huchra J.P., Davis M., Yahil A., Fisher K.B. and Tonry J. (1992) A redshift
survey of IRAS galaxies VII. The infrared and redshift data for the 1.936 Jansky
sample. Astrophys. J. Suppl. Ser. 83, 2963.
Strauss M.A. and Willick J.A. (1995) The density and peculiar velocity fields of nearby
galaxies. Phys. Reports 261, 271431.
Sutherland W. and Efstathiou G. (1991) Correlation functions of rich clusters of galaxies.
Mon. Not. R. Astr. Soc. 248, 159167.
Sylos Labini F. and Montuori M. (1998) Scale invariant properties of the APM-Stromlo
survey. Astr. Astrophys. 331, 809814.
Sylos Labini F., Montuori M. and Pietronero L. (1998) Scale-invariance of galaxy clustering. Phys. Rep. 293, 61226.
Szalay A.S., Matsubara T. and Landy S.D. (1998) Redshift space distortions of the correlation function in wide angle galaxy surveys. Astrophys. J. Lett. 498, L1L4.
Szapudi I. (1998) A new method for calculating counts in cells. Astrophys. J. 497, 1620.
Szapudi I. (2000) Cosmic statistics of statistics: N -point correlations. http://arXiv.org/abs/
astro-ph/0008224.
Szapudi I., Branchini E., Frenk C.S., Maddox S. and Saunders W. (2000) The luminosity
dependence of clustering and higher order correlations in the PSCz survey. Mon. Not.
R. Astr. Soc. 318, L45L50.
Szapudi I. and Colombi S. (1996) Cosmic error and statistics of large-scale structure.
Astrophys. J. 470, 131148.
Szapudi I., Colombi S. and Bernardeau F. (1999) Cosmic statistics of statistics. Mon. Not.
R. Astr. Soc. 310, 428444.
Szapudi I. and Szalay A.S. (1993) Higher order statistics of the galaxy distribution using
generating functions. Astrophys. J. 408, 4356.
Szapudi I. and Szalay A.S. (1998) A new class of estimators for the N -point correlations.
Astrophys. J. Lett. 494, L41L44.
Tadros H., Ballinger W.E., Taylor A.N., Heavens A.F., Efstathiou G., Saunders W., Frenk
C.S., Keeble O., McMahon R., Maddox S.J., Oliver S., Rowan-Robinson M., Sutherland W.J. and White S.D.M. (1999) Spherical harmonic analysis of the PSCz galaxy
catalogue: Redshift distortions and the real-space power spectrum. Mon. Not. R. Astr.
Soc. 305, 527546.
Takayasu H. (1989) Fractals in the Physical Sciences. Manchester University Press,
Manchester.
Taylor A. and Valentine H. (1999) The inverse redshift-space operator: Reconstructing
cosmological density and velocity fields. Mon. Not. R. Astr. Soc. 306, 491503.
Tegmark M. (1995) A method for extracting maximum resolution power spectra from
galaxy surveys. Astrophys. J. 455, 429438.
Tegmark M. (1997) How to measure CMB power spectra without losing information. Phys.
Review D 55, 58955907.
428
2002 by CRC Press LLC
References
419
Tegmark M., Hamilton A.J.S., Strauss M.A., Vogeley M.S. and Szalay A.S. (1998) Measuring the galaxy power spectrum with future redshift surveys. Astrophys. J. 499, 555576.
Tegmark M., Taylor A.N. and Heavens A.F. (1997) KarhunenLo`ewe eigenvalue problems
in cosmology: How should we tackle large data sets? Astrophys. J. 480, 2235.
Tegmark M., Zaldarriaga M. and Hamilton A.J.S. (2001) Towards a refined cosmic concordance model: Joint 11-parameter constraints from CMB and large-scale structure. Phys.
Rev. D 63, 043007.
Toomet O., Andernach H., Einasto J., Einasto M., Kasak E., Starobinsky A.A. and Tago
E. (1999) The superclustervoid network V. Alternative evidence for its regularity.
http://arXiv.org/abs/astro-ph/9907238.
Tucker D.L., Oemler A., Kirshner R.P., Lin H., Shectman S.A., Landy S.D., Schechter
P.L., Muller V., Gottlober S. and Einasto J. (1997) The Las Campanas redshift survey
galaxy-galaxy autocorrelation function. Mon. Not. R. Astr. Soc. 285, L5L9.
Tully R.B. and Fisher J.R. (1977) A new method of determining distances to galaxies.
Astron. Astrophys. 54, 661673.
Turner E.L. and Gott J.R. (1976) Groups of galaxies. I. A catalog. Astrophys. J. Suppl. Ser.
32, 409427.
Upton G. and Fingleton B. (1985) Spatial Data Analysis by Example, vol. 1. John Wiley &
Sons, Chichester.
Valentine H., Saunders W. and Taylor A. (2000) Reconstructing PSCz with a generalized
PIZA. Mon. Not. R. Astr. Soc. 319, L13L17.
van de Weygaert R. (1991) Voids and the Geometry of Large Scale Structure. Ph.D. thesis,
University of Leiden, Leiden.
van de Weygaert R. (1994) Fragmenting the Universe. 3: The constructions and statistics
of 3-D Voronoi tessellations. Astron. Astrophys. 283, 361406.
van de Weygaert R. and Bertschinger E. (1996) Peak and gravity constraints in Gaussian
primordial density fields: An application of the HoffmanRibak method. Mon. Not. R.
Astr. Soc. 281, 84118.
van de Weygaert R. and Hoffman Y. (2000) The structure of the local universe and the
coldness of the cosmic flow. In Cosmic Flows Workshop, ASP Conference Series (eds.
S. Courteau and J. Willick), vol. 201, Astronomical Society of the Pacific, San Francisco, pp. 169176.
van de Weygaert R. and Icke V. (1989) Fragmenting the universe II. Voronoi vertices as
Abell clusters. Astron. Astrophys. 213, 19.
van de Weygaert R., Jones B.J.T. and Martnez V.J. (1992) The minimal spanning tree as
an estimator for generalized dimensions. Phys. Lett. A 169, 145150.
van Haarlem M. (1996) Projection effects in the Abell catalogue. In Mapping, Measuring, and Modelling the Universe, ASP Conference Series (eds. P. Coles, V.J. Martnez
and M.J. Pons-Bordera), vol. 94, Astronomical Society of the Pacific, San Francisco,
pp. 191196.
van Lieshout M.N.M. and Baddeley A. (1996) A non-parametric measure of spatial interaction in point patterns. Stat. Neerlandica 50, 344361.
van Waerbeke L., Mellier Y., Radovich M., Bertin E., Dantel-Fort M., McCracken H.J.,
F`evre O.L., Foucaud S., Cuillandre J.C., Erben T., Jain B., Schneider P., Bernardeau F.
and Fort B. (2001) Cosmic shear statistics and cosmology. Astr. Astrophys. 374, 757
769.
429
2002 by CRC Press LLC
420
References
Verde L., Heavens A.F., Matarrese S. and Moscardini L. (1998) Large-scale bias in the
Universe II: redshift space bispectrum. Mon. Not. R. Astr. Soc. 300, 747757.
Verde L., Wang L., Heavens A.F. and Kamionkowski M. (2000) Large-scale structure, the
cosmic microwave background, and primordial non-Gaussianity. Mon. Not. R. Astr. Soc.
313, 141145.
Viana P.T.P. and Liddle A.R. (1999) Galaxy clusters at 0:3 < z < 0:4 and the value of
0 .
Mon. Not. R. Astr. Soc. 303, 535546.
Vogeley M.S., Park C., Geller M.J., Huchra J.P. and Gott J.R.I. (1994) Topological analysis
of the CfA redshift survey. Astrophys. J. 420, 525544.
Vogeley M.S. and Szalay A.S. (1996) Eigenmode analysis of galaxy redshift surveys.
I. Theory and methods. Astrophys. J. 465, 3453.
Voronoi G. (1908) Nouvelles applications des parametres continus a la theorie des fromes
quadratiques. J. Reine Angew. Math. 134, 198287.
Walker J.S. (1998) A Primer on Wavelets and Their Scientific Applications. Chapman &
Hall, Boca Raton.
Wand M.P. and Jones M.C. (1995) Kernel Smoothing. Chapman & Hall, London.
Wang Y.H. (2000) Fiducial intervals: what are they? Amer. Statistician 54, 105111.
Walder U. and Stoyan D. (1996) On variograms and point process statistics. Biom. J. 38,
895905.
Webb S. (1999) Measuring the Universe. The Cosmological Distance Ladder. SpringerVerlag in association with Praxis Publishing, Chichester.
Weinberg D.H., Gott J.R.I. and Melott A.L. (1987) The topology of large-scale structure
I. Topology and the random phase hypothesis. Astrophys. J. 321, 227.
White M. (2001) The redshift space power spectrum in the halo model. Mon. Not. R. Astr.
Soc. 321, 13.
White S.D.M. (1979) The hierarchy of correlation functions and its relation to other measures of galaxy clustering. Mon. Not. R. Astr. Soc. 186, 145154.
Willick J.A., Courteau S., Faber S.M., Burstein D., Dekel A. and Strauss M.A. (1997)
Homogeneous velocitydistance data for peculiar velocity analysis. III. The Mark III
catalog of galaxy peculiar velocities. Astrophys. J. Suppl. 109, 333366.
Willick J.A., Strauss M., Dekel A. and Kolatt T. (1997) Maximum likelihood comparisons
of TullyFisher and redshift data: Constraints on
and biasing. Astrophys. J. 486, 629
664.
Witten T.A. and Sander L.M. (1981) Diffusion-limited aggregation, a kinetic critical phenomenon. Phys. Rev. Lett. 47, 14001403.
Wu K.K.S., Lahav O. and Rees M.J. (1999) The large-scale smoothness of the Universe.
Nature 397, 225235.
Yahil A., Strauss M.A., Davis M. and Huchra J.P. (1991) A redshift survey of IRAS galaxies. II. Methods for determining self-consistent velocity and density fields. Astrophys. J.
372, 380393.
Yoshida N., Colberg J.M., White S.D.M., Evrard A.E., MacFarland T.J., Couchman
H.M.P., Jenkins A., Frenk C.S., Pearce F.R., Efstathiou G., Peacock J.A. and Thomas
P.A. (2001) Simulation of deep pencil-beam redshift surveys. Mon. Not. R. Astr. Soc.
325, 803816.
Zaroubi S., Hoffman Y. and Dekel A. (1999) Wiener reconstruction of large-scale structure
from peculiar velocities. Astrophys. J. 520, 413425.
430
2002 by CRC Press LLC
References
421
Zaroubi S., Hoffman Y., Fisher K.B. and Lahav O. (1995) Wiener reconstruction of the
large scale structure. Astrophys. J. 449, 446459.
Zeldovich Y.B. (1970) Gravitational instability: An approximate theory for large density
perturbations. Astr. Astrophys. 5, 8489.
Ziman J.M. (1979) Models of Disorder: The Theoretical Physics of Homogeneously Disordered Systems. Cambridge University Press, Cambridge.
Zwicky F., Herzog E. and Wild P. (19611968) Catalogue of Galaxies and of Clusters of
Galaxies. California Institute of Technology (CIT), Pasadena.
431
2002 by CRC Press LLC
432
2002 by CRC Press LLC
424
433
2002 by CRC Press LLC