You are on page 1of 183

MODERN COSMOLOGY

The Universe in its Evolution and Structure


Max Camenzind
Landessternwarte K onigstuhl Heidelberg
July 19, 2009
The big bang appears to the casual observer as just another myth, albeit
without some of the more obvious anthropocentric characteristics. The
difference, however, is that modern cosmology is based upon the scientic
method. The scientic method has very specic rules.
Preface
In the history of science few developments have been more important than the advent of the
new heliocentric cosmology in the sixteenth and seventeenth centuries. Whereas most of the
ancient Greeks and European medievals had believed the earth was at the center of the universe
- with the sun, moon, planets, and stars orbiting around us - in the sixteenth century a new
idea began to emerge. According to this new way of thinking, it was not the earth but the sun
that was at the center of the cosmic system. This revolutionary idea was famously proposed
by Nicholas Copernicus in his book On the Revolution of the Heavenly Spheres, published in
1543.
The progress of modern cosmology has been guided by both observational and theoretical
advances. The subject really took off in 1917 with a paper by Albert Einstein that attempted
the ambitious task of describing the universe by means of a simplied mathematical model.
Five years later Alexander Friedmann constructed models of the expanding universe that had
their origin in a big bang. These theoretical investigations were followed in 1929 by the pi-
oneering work on nebular redshifts by Edwin Hubble and Milton Humason, who provided
the observational foundations of presentday cosmology. In 1948 the steady state theory of
Hermann Bondi, Thomas Gold, and Fred Hoye added a spice of controversy that led to many
observational tests, essential for the healthy growth of the subject as a branch of science. Then
in 1965 Arno Penzias and Robert Wilson discovered the microwave background, which not
only revived George Gamows concept of the hot big bang proposed nearly two decades be-
fore, but also prompted even more daring speculations about the early history of the universe.
Present models of the universe hold two fundamental premises: the cosmological principle
and the dominant role of gravitation. Derived by Hubble, the cosmological principle holds
that if a large enough sample of galaxies is considered, the universe looks the same from all
positions and in all directions in space. The second point of agreement is that gravitation
is the most important force in shaping the universe. According to Einsteins general theory
of relativity, which is a geometric interpretation of gravitation, matter produces gravitational
effects by actually distorting the space about it; the curvature of space is described by a form
of non-Euclidean geometry. A number of cosmological theories satisfy both the cosmological
principle and general relativity. The two main theories are the big-bang hypothesis and the
steady-state hypothesis, with many variations on each basic approach.
The material here is therefore of two kinds: relevant pieces of physics and astronomy that
are often not found in undergraduate courses, and applications of these methods to more recent
research results. The former category is dominated by general relativity and quantum elds.
Relativistic gravitation has always been important in the large-scale issues of cosmology, but
the application of modern particle physics to the very early universe is a more recent develop-
ment. Many excellent texts exist on these subjects, but I want to focus on those aspects that
are particularly important in cosmology. At times, I have digressed into topics that are strictly
unnecessary, but which were just too interesting to ignore. These are, after all, the crown
iv Preface
jewels of 20th Century physics, and I rmly believe that their main features should be a stan-
dard part of a graduate education in physics and astronomy. Despite this selective approach,
which aims to concentrate of the essential core of the subject, the course will cover a wide
range of topics. My original plan was in fact focused more specically on matters to do with
particle physics, the early universe, and structure formation. However, as I wrote, the subject
imposed its own logic: it became clear that additional topics simply had to be added in order
to tell a consistent story. This tendency for different parts of cosmology to reveal unexpected
connections is one of the joys of the subject, and is also a mark of the maturity of the eld.
According to big-bang theories, at the beginning of time, all of the matter and energy in
the universe was concentrated in a very dense state, from which it exploded, with the resulting
expansion continuing until the present. This big bang is dated between 12 and 15 billion years
ago. In this initial state, the universe was very hot and contained a thermal soup of quarks,
electrons, photons, and other elementary particles. The temperature rapidly decreased, falling
from 10
13
degrees Kelvin after the rst microsecond to about one billion degrees after three
minutes. As the universe cooled, the quarks condensed into protons and neutrons, the building
blocks of atomic nuclei. Some of these were converted into helium nuclei by fusion; the
relative abundance of hydrogen and helium is used as a test of the theory. After many millions
of years the expanding universe, at rst a very hot gas, thinned and cooled enough to condense
into individual stars and galaxies, and even Black Holes.
Several spectacular discoveries since 1960 have shed new light on the problem. Optical
and radio astronomy complemented each other in the discovery of the quasars and the radio
galaxies. It is believed that the energy reaching us now from some of these objects was emitted
not long after the creation of the universe. Further evidence for the big-bang theory was the
discovery in 1965 that a cosmic background noise is received from every part of the sky. This
background radiation has the same intensity and distribution of frequencies in all directions
and is not associated with any individual celestial object. This radiation lling space has a
black body temperature of 2.73 K and is interpreted as the electromagnetic remnant of the
primordial reball, stretched to long wavelengths by the expansion of the universe. More
recently, the analysis of radiation from distant celestial objects detected by articial satellites
has given additional evidence for the big-bang theory and the small scale structure imprinted
by early uctuations in the density distribution.
Max Camenzind
Heidelberg, July 17, 2009
Contents
I The Observable Universe 1
1 History and Overview 3
1.1 History of the Universe Myths and Science . . . . . . . . . . . . . . . . . 3
1.2 Steps towards Modern Cosmology . . . . . . . . . . . . . . . . . . . . . . . 6
1.3 Key Facts of Modern Cosmology . . . . . . . . . . . . . . . . . . . . . . . . 7
1.4 Problems and Mysteries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.5 The Edge of Time . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.6 Textbooks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.7 Units . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2 The Observable Universe 13
2.1 The Universe is Expanding and Accelerating . . . . . . . . . . . . . . . . . . 15
2.2 Cosmic Background Radiation as a Relic of the Early Universe . . . . . . . . 22
2.3 Galaxies are not Uniformly Distributed . . . . . . . . . . . . . . . . . . . . 30
2.4 Gravity is Dominated by Dark Matter . . . . . . . . . . . . . . . . . . . . . 33
2.5 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
II Relativistic World Models 43
3 The Relativistic Cosmos 45
3.1 Relativity and Cosmology . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
3.2 Einsteins Vision of Gravity . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
3.2.1 The Concept of SpaceTime . . . . . . . . . . . . . . . . . . . . . . . 47
3.2.2 Gravity is an Afne Connection on SpaceTime . . . . . . . . . . . . 51
3.2.3 Differential Forms on SpaceTime . . . . . . . . . . . . . . . . . . . 55
3.2.4 Curvature of SpaceTime . . . . . . . . . . . . . . . . . . . . . . . . 57
3.2.5 Curvature and Einsteins Equations . . . . . . . . . . . . . . . . . . 58
3.3 General Relativity is the Correct Theory of Gravity . . . . . . . . . . . . . . 62
3.4 Isotropic SpaceTimes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
3.4.1 Slicing of SpaceTime . . . . . . . . . . . . . . . . . . . . . . . . . . 65
3.4.2 Isotropic Riemannian Spaces . . . . . . . . . . . . . . . . . . . . . . 68
3.4.3 Spaces of Constant Curvature . . . . . . . . . . . . . . . . . . . . . 69
3.4.4 FriedmannRobertsonWalker (FRW) SpaceTimes . . . . . . . . . . 70
vi Contents
3.5 The FRW World Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
3.6 The Origin of the Cosmological Redshift . . . . . . . . . . . . . . . . . . . . 73
3.7 The Luminosity Distance and the HubbleLaw . . . . . . . . . . . . . . . . 74
3.8 The Hubble Constant . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
3.9 The Apparent Angular Width of Galaxies . . . . . . . . . . . . . . . . . . . 79
3.10 Number Counts in the Expanding Universe . . . . . . . . . . . . . . . . . . 81
3.11 Slightly Inhomogeneous World Models . . . . . . . . . . . . . . . . . . . . 83
3.12 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
3.13 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
4 The Universe with Matter and Dark Energy 87
4.1 Description of Matter in a Relativistic Cosmos . . . . . . . . . . . . . . . . . 87
4.1.1 Fluid Approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
4.1.2 Kinetic Description . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
4.1.3 EOS for Vacuum Energy . . . . . . . . . . . . . . . . . . . . . . . . 91
4.2 Einsteins Equations for FRW Models . . . . . . . . . . . . . . . . . . . . . 92
4.2.1 Derivation of Friedmanns Equations . . . . . . . . . . . . . . . . . 92
4.2.2 Energy Conservation . . . . . . . . . . . . . . . . . . . . . . . . . . 95
4.2.3 Density Parameters for Friedman Models . . . . . . . . . . . . . . . 96
4.3 FRW Models without Vacuum Energy . . . . . . . . . . . . . . . . . . . . . 97
4.3.1 The Euclidean Universe, k=0 . . . . . . . . . . . . . . . . . . . . . . 98
4.3.2 The Closed Universe, k=1 . . . . . . . . . . . . . . . . . . . . . . . 98
4.3.3 The Open Universe, k=1 . . . . . . . . . . . . . . . . . . . . . . . 99
4.3.4 The Singularity at t=0: BigBang . . . . . . . . . . . . . . . . . . . 100
4.3.5 The Mattig Formula for the Luminosity Distance . . . . . . . . . . . 100
4.4 The Present Universe with Dark Energy . . . . . . . . . . . . . . . . . . . . 103
4.4.1 Cosmological Parameters . . . . . . . . . . . . . . . . . . . . . . . . 104
4.4.2 Solutions for Inationary Universes . . . . . . . . . . . . . . . . . . 105
4.4.3 Age of the Universe . . . . . . . . . . . . . . . . . . . . . . . . . . 105
4.4.4 The Event Horizon . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
4.4.5 The Particle Horizon . . . . . . . . . . . . . . . . . . . . . . . . . . 111
4.4.6 Conformal Maps of the Present Universe . . . . . . . . . . . . . . . 112
4.5 Measuring the Acceleration of the Present Universe . . . . . . . . . . . . . . 122
4.5.1 Luminosity Distance and HubbleDiagrams for CDM . . . . . . . 123
4.5.2 Measuring Cosmology with Supernovae . . . . . . . . . . . . . . . . 129
4.6 Angular Width in FRW Models . . . . . . . . . . . . . . . . . . . . . . . . . 147
4.7 Redshift Distribution of Cosmological Objects . . . . . . . . . . . . . . . . . 149
4.8 The Cosmological Fundamental Plane . . . . . . . . . . . . . . . . . . . . . 153
4.9 On the Origin of the Dark Energy . . . . . . . . . . . . . . . . . . . . . . . 154
4.9.1 The Vacuum is not Empty . . . . . . . . . . . . . . . . . . . . . . . 156
4.9.2 Quintessence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156
4.9.3 Braneworld Models Higher Dimensions . . . . . . . . . . . . . . . 159
4.9.4 Effects from Inhomogeneous Universe . . . . . . . . . . . . . . . . . 166
4.10 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168
Contents vii
List of Figures 171
List of Tables 173
Part I
The Observable Universe
1 History and Overview
Our exploration of cosmology begins with a brief history of the human desire to understand the
cosmos. Mythology, humanitys rst attempt to grapple with cosmological questions, consists
of narrative tales that describe the universe in understandable terms. The text briey discusses
the Tanzanian myth The Word as an example of a creation myth. Cosmology, particularly as
expressed by a mythology, can inuence a cultures or an individuals actions.
1
1.1 History of the Universe Myths and Science
The big bang appears to the casual observer as just another myth, albeit without some of the
more obvious anthropocentric characteristics. The difference, however, is that modern cos-
mology is based upon the scientic method. The scientic method has very specic rules. It
is based on objective data, observations that are independent of who made those observations.
Once sufcient data are collected, a hypothesis is framed to explain and unify them. In order
to be regarded as scientic, the hypothesis must meet at minimum ve characteristics: it must
be relevant, testable, consistent, simple, and have explanatory power. Of these, the quality
of testability particularly denes the scientic method. A hypothesis that does not contain
the potential to be falsied is not scientic. Once a hypothesis has met success at explain-
ing data and has proven itself useful in predicting new phenomena, it is generally called a
theory. Some particularly well established theories, especially those pertaining to a limited
phenomenon or forming the foundation for a broader theory, are called laws. Hence we refer
to the law of gravity, even though scientic laws are subject to modication as our under-
standing improves. Newtons law of gravity is extremely well veried for the regime in which
it is applicable (weak gravitational elds and speeds small compared to the speed of light);
but Newtons law must be superseded by Einsteins general theory of relativity.
The rst attempt to construct a systematic cosmology that was grounded in physical theory
was the model of Aristotle. Aristotle developed a theory of motion, and dened the concepts
of natural motion and force. In Aristotles view, the Earth was the center of the universe
and the center of all natural motions. Motions on the Earth were linear and nite, while the
heavenly bodies executed perfect circles forever. The stars and planets were composed of a
perfect element called ether, whereas Earthly objects were made up of varying combinations
of the four ancient elements of earth, air, re and water; a bodys motion was a consequence
of its composition. Although our modern denitions of these concepts are quite different from
Aristotles, natural motion and force remain fundamental to our understanding of the structure
and evolution of the universe. Aristotles Earth-centered worldview (Fig. 1.1) was embodied
in the detailed model of Ptolemy, with its deferents, epicycles, and eccentrics, all designed
1
We follow here John F. Hawleys book on Foundations of Modern Cosmolgy, Oxford University Press
4 1 History and Overview
to predict the complicated celestial motions of the planets while still requiring motion in the
heavens to be built upon circles.
Figure 1.1: Aristotles crystalline sphere cosmos is the rst cosmological model. The illustration derives
from the Cosmographia seu descriptio totius orbis (1524) of Petrus Apianus (Peter Apian, 1495-1552).
Earth is at the center of the cosmos, xed and stable, followed by the Moon, Mercury, Venus, the
Sun, Mars, Jupiter, Saturn, and nally, three medieval ether regions that culminated with the Empyrean
Sphere. The ordering of the planets differed, so far as we know, between Aristotle and Ptolemy. There
was much debate throughout antiquity and the middle ages about the precise relation of the planets and
the Sun. The number of spheres given by Eudoxus (ca. 390 B.C.) is 27 (including the sphere of the Fixed
Stars). Callippus (ca. 370 B.C.) increases this to 34. Aristotle (384-322 BC) improves on Callippus by
including additional spheres to counteract some of the motions of the planetary spheres. These additional
spheres are placed between the outermost sphere of a given planet and the innermost sphere of the next
planet and are one less than the number of spheres of the latter.
Aristotle (384-322 BC) surveyed the whole of human knowledge of his time, and wrote a
lot of works on philosophy and science. Most of the surviving works are not from the many
volumes Aristotle wrote personally but are thought to be lecture notes, perhaps recorded by
a student at the Lyceum. Thus it may not be too surprising that many of them are difcult to
follow and sometimes even self-contradictory! Aristotles cosmos is divided into the perfect,
unchanging heavens, and the imperfect sublunary sphere (everything on Earth, within the
orbit of the Moon). Everything was thought to have a natural place within this scheme, and
natural motion was the result of objects natural tendency to attain their places. The passage
quoted here begins with Aristotles reasons for believing in a fth element, sometimes called
1.1 History of the Universe Myths and Science 5
quintessence (watch for this to show up much later in the course!) or ether, apart from the
usual four (earth, re, air, and water). He argues that bodies have natural motions (as opposed
to unnatural ones). He thinks re and air naturally move upwards in a line, while earth and
water move down. But he argues that there is a simple body (or element) for each simple
motion, and circular motion is also a simple motion, so there must be a new element whose
motion is naturally circular.
During the Renaissance, humanitys cosmological model changed dramatically. The rst
blow in the Scientic Revolution was struck by Copernicus, whose Sun-centered model of
the heavens gained rapid ascendency in Renaissance Europe. Tycho Brahes detailed naked
eye observations of the heavens provided the data that Kepler used to derive his laws of plan-
etary motion. Keplers laws of planetary made it possible for the rst time for humans to
understand the paths of the wanderers across the sky.
A consequence of Keplers second law is that planets orbit more slowly the more distant
they are from the Sun. The third law enables the period of a planet, comet, or asteroid to be
computed once observations establish the length of the semimajor axis of its orbit. These laws
were among the greatest quantitative achievements of the Renaissance.
Kepler and Galileo were contemporaries, though Kepler was more of a theorist and Galileo
was primarily an observer. Galileo was the rst to make serious scientic use of the telescope,
an instrument which provided observations that challenged the Ptolemaic model of the heav-
ens. (Kepler was unable to afford to purchase a telescope, a prohibitively expensive device at
the time, though he was able to borrow one for a summer from a visiting nobleman. Galileo
promised for several years to make a telescope for Kepler, but never got around to fullling
his promise.) Galileo observed craters on the Moon, demonstrating that it was not a perfect,
smooth sphere; he also gave the large lunar plains the name of maria (seas) because he
thought they might be lled with water. He also found that the Milky Way was not a solid
band of light but was lled with myriad stars, too small to be resolved by the unaided eye.
Another key observation by Galileo was that Venus went through a full cycle of phases, just
like the Moon; this was impossible in the Ptolemaic model but was required by the Coperni-
can model, since Venus is between the Earth and the Sun in the latter. But one of Galileos
most important discoveries was of the four largest satellites of Jupiter, now called the Galilean
moons. These bodies demonstrated that the Earth was not the only center of motion in the
universe, thus refuting one of the important tenets of Ptolemaic-Aristotelian cosmology and
physics.
Galileo also studied mechanics. From direct observation and careful reasoning, he was
able to arrive at the conclusion that all bodies fall at the same rate, if air resistance is negligible.
This principle, now called the equivalence principle, is one of the foundations of the general
theory of relativity, though Galileo could not have appreciated this at the time. Galileo also
realized that motion might not be easily detectable by observers partaking of that motion, i.e.
that motion is relative, though he never succeeded in working out the laws of motion. But a
few months after Galileos death, Isaac Newton was born on a farm in Lincolnshire, England,
beginning a life that would complete the Copernican Revolution with the fundamental laws of
physics and gravitation that govern the universe under most conditions.
The Aristotelian cosmological system was consistent with his physics. Although Aristo-
tles cosmology is quite different from the modern point of view, in what ways is it consistent
with modern ideas? How did it differ? What were some of the motivations for Copernicus to
propose such a grand change to the prevailing concepts of the universe?
Tycho Brahes observations of a supernova that appeared in 1572 helped to end the belief
6 1 History and Overview
in the Aristotelean unchanging perfection of the celestial realm. How would you feel if you
observed something that so challenged a basic tenet of your world view? In Tychos own
words:
Amazed, and as if astonished and stupeed, I stood still, gazing for a certain length of
time with my eyes xed intently upon it and noticing that same star placed close to the stars
which antiquity attributed to Cassiopeia. When I had satised myself that no star of that kind
had ever shone forth before, I was led into such perplexity by the unbelievability of the thing
that I began to doubt the faith of my own eyes.
Kepler too was to observe a supernova, only 32 years later in 1604. The next supernova
visible to the naked eye did not occur until 1987 when a star exploded in the nearby irregular
galaxy known as the Large Magellanic Cloud.
1.2 Steps towards Modern Cosmology
Controversy over the nature of spiral nebulae had persisted since the late 18th century, with
one camp insisting they were external universes, while their opponents were equally con-
vinced that the spiral nebulae were localized clusters of stars within our Galaxy. An important
early discovery was Shapleys determination of the size of the Milky Way Galaxy, and of our
location within it. Shapley found the Milky Way to be much larger than previously believed,
and on this basis he erroneously concluded that the spiral nebulae must be relatively nearby
clusters. Shapley and Curtis participated in a famous debate in 1920 over the nature of the
spiral nebulae, but insufcient data prevented a resolution of the puzzle. Finally, Hubble de-
termined that the Andromeda Nebula (now known as the Andromeda Galaxy) was much too
far to lie within the connes of the Milky Way; Hubble had discovered external galaxies. In
the rst quarter of the twentieth century, humanitys view of the cosmos leaped from a fairly
limited realm of the Sun surrounded by an amorphous grouping of stars, to one in which the
Milky Way is just a typical spiral galaxy in a vast universe lled with galaxies.
Not long after Hubbles discovery of external galaxies came his discovery of a linear rela-
tionship between their redshifts and their distances, a relationship known today as the Hubble
Law. Determining the value of the constant of proportionality, the Hubble constant, remains
an important research goal of modern astronomy. The Hubble constant is not really constant,
because it can change with time, though at any given instant of cosmic time in a homogeneous,
isotropic universe, it is the same at all spatial locations. The inverse of the Hubble constant,
called the Hubble time, gives an estimate of the age of the universe.
The development of the theory of general relativity provided the framework in which Hub-
bles discovery could be understood. Einstein found that his equations would not admit a
static, stable model of the universe, even with the addition of the cosmological constant.
The timely discovery of the redshift-distance relationship provided evidence that the universe
was not static, but was expanding. The Robertson-Walker metric is the most general metric
for an isotropic, homogeneous universe that is also dynamic; i.e. it changes with time. An im-
portant parameter in this metric is the scale factor, the quantity which describes how lengths in
the universe change with cosmic time. The scale factor can be used to derive the cosmological
redshift, the change in wavelength of light as it traverses the universe.
Measuring Hubbles constant requires accurate distances to increasingly remote galaxies.
One of the best distance measures is the Cepheid variable star. The HST has now been able to
detect Cepheid variable stars in the galaxy M100 in the Virgo galaxy cluster. Several Cepheids
1.3 Key Facts of Modern Cosmology 7
have been found such as this one. These new data give us a distance to M100 of 17 Mpc and
is consistent with a rather large Hubble constant of about 70 km/sec/Mpc. This value has now
been conrmed by Supernovae type Ia measurments and a detailed analysis of the cosmic
microwave anisotropies by the WMAP satellite. It is the rst time in the history of the Hubble
constant that a converging value has been acchieved by various methods.
1.3 Key Facts of Modern Cosmology
The relativistic cosmos is expanding. The history of the universe can be divided into many
eras, depending on which constituent was most important. Today the matter density of the uni-
verse dominates its gravity completely; thus we say that we live in the matter era. However,
conditions were not always as they are today. The relationship between temperature and red-
shift of the CMBR demonstrates that as we look backward in time toward t = 0, the photons
in the CMBR become increasingly hot. Density also changes with time, becoming ever higher
as we look toward earlier times; but the variation with scale factor is different for matter den-
sity and for radiation energy density, with matter varying as the cube of the scale factor, while
radiation energy density varies as the fourth power of R. Hence there was a time in the past
when radiation was more important than ordinary matter in determining the evolution of the
universe. This high-temperature epoch denes the early universe. Temperature is a measure
of energy, and Einsteins equation E = mc
2
tells us that energy and matter are equivalent.
At sufciently high temperatures, particles with large mass can be created, along with their
antiparticles, from pure energy. The temperature also inuenced how the fundamental forces
of nature behaved during the earliest intervals of the universes history.
The interval of domination by radiation is called the radiation era. As we approach t=0
in this era, we encounter increasingly unfamiliar epochs, dominated by different physics and
different particles. The earliest was the Planck epoch, during which all four fundamental
forces were unied and particles as we know them could not have existed. Next followed
the GUT epoch, when gravity had decoupled but the other three forces remained unied.
The small excess of matter that makes up the universe today must have been created during
this epoch, by a process still not completely understood. As the temperature dropped, the
universe traversed the quark epoch, the hadron epoch, the lepton epoch, and the epoch of
nucleosynthesis.
The rst atomic nuclei formed during the nucleosynthesis epoch. Most of the helium in the
universe was created from the primordial neutrons and protons by the time the nucleosynthesis
epoch ended scarcely three minutes after the big bang. A few other trace isotopes, specically
deuterium (heavy hydrogen) and lithium 7, were also created, and their density depends sen-
sitively upon the density of the universe during this time. If the universe were too dense,
then most of the deuterium would have fused into helium. Only in a low-density universe
can the deuterium survive. The major factor controlling the ultimate densities of helium and
deuterium is the abundance of neutrons. The more neutrons that decay before combining with
protons, the smaller the abundances of heavier elements. The availability of neutrons depends
on the expansion rate as well as the cosmic matter density. Comparing the observed densities
of the primordial isotopes to those computed from models and translating the results into
B
,
the density parameter, gives
B
= 0.015/h
2
where h is the Hubble parameter divided by
100 km/sec/Mpc. The smaller H
0
, the larger ; if H
0
=50,
B
is approximately 0.06, whereas
H
0
= 100 gives
B
of only 0.015. This range is still much less than = 1, but nucleosynthe-
8 1 History and Overview
sis limits can indicate only the density of baryons, because only baryons participate in nuclear
reactions. Hence we must conclude that the universe contains less than the critical density of
baryons.
After nucleosynthesis, nothing much happened for roughly a million years as the universe
continued to cool. The ordinary matter consisted of a hot plasma of nuclei and electrons. The
free electrons made the plasma opaque; a photon of radiation could not have traveled far before
being scattered. However, once the universe cooled to approximately 3000 K, the electrons no
longer moved fast enough to escape the attraction of the nuclei, and atoms formed. Although
there had been no previous combination, this event is still known as recombination. The last
moment at which the universe was opaque forms the surface of last scattering; it represents
the effective edge of the universe that is even theoretically visible with an optical telescope,
since no optical telescope could ever penetrate the dense, opaque plasma that existed prior to
recombination. Once the radiation was able to stream freely through the universe, matter and
radiation lost the tight coupling that had bound them since the beginning. Henceforth matter
and radiation evolved almost entirely independently. The photons that lled the universe at
the surface of last scattering make up the CMBR today, but now their energy is mostly in
the microwave band. At some point before or near recombination, the matter density and
the energy density were equally important. This is the epoch in which structure formation
began to occur. The seeds of structure formation may have been planted much earlier, during
the GUT epoch, but the tight coupling between radiation and matter prevented the density
perturbations from doing much. Once matter and radiation went their separate ways, density
pertubations could evolve on their own. The most overdense areas collapsed gravitationally,
forming galaxies and clusters of galaxies. Less dense areas probably led to voids, the large
underdense areas we see on the sky today. The process by which structure formed in the early
universe is still very poorly understood; better data from instruments such as the planned
successors to COBE (WMAP and Planck) will help to elucidate the mystery of the galaxies.
The matter density is one of the fundamental parameters of the universe; in the standard
( = 0) models, the matter density determines the geometry of the cosmos. An accurate
determination of this quantity is thus of great cosmological importance. Many methods have
been developed to try to measure the matter density. Most rely upon detecting the orbits of
visible matter and using Keplers laws to compute the mass; these dynamical methods are
probably the most widely employed. At all scales, we have found that the amount of matter
revealed through gravitational interactions is greater than can be explained by the mass of the
visible stars. The nature of this unseen dark matter is one of the most important outstanding
cosmological problems.
At the scale of individual galaxy disks, much of the dark matter can be attributed to stellar
ashes such as white dwarfs, neutron stars, and black holes, as well as to extremely faint objects
such as very low-mass stars and brown dwarfs. However, at larger scales, the total aggregation
of known baryonic matter cannot account for the mass that orchestrates gravitational interac-
tions. Galaxies, including the Milky Way, appear to be surrounded by huge spheroidal dark
halos. The composition of the dark halos is unknown, but observations have indicated that
at least some of the mass must take the form of compact objects called MACHOs. Since we
believe that only baryonic (i.e. ordinary) matter can form compact objects, this suggests that
galaxies are surrounded by an invisible cloud of objects such as neutron stars or small black
holes. How and when these bodies formed is still a mystery.
At larger scales, measurements of rise until they typically level off within a range of
0.1 to 0.3, with the middle of this range (about 0.25) currently seeming to be most likely.
1.4 Problems and Mysteries 9
This is far greater than can be accommodated by the abundances of primordial elements such
as helium and deuterium. Thus we conclude that some 90% of the matter of the universe
is not only invisible, but is nonbaryonic. If it is not baryons, what is it? It must consist
of some type of WIMP (weakly interacting massive particle). What this particle might be
remains unknown, though particle physicists can provide plenty of possibilities. One obvious
candidate is the neutrino. Evidence has grown over the last decade that the neutrino actually
has an extremely tiny mass. There is no theoretical reason that the neutrino must be massless;
indeed, one hypothesis for the dearth of solar neutrinos observed over the past 30 years is
that neutrinos are massive, and the kind that can be observed is converted into a kind that
is invisible to most neutrino detectors on the journey from the Sun. Laboratory evidence
is beginning to support this hypothesis. If neutrinos have even a small mass per particle,
they could add considerably to the cosmic matter density because they are approximately as
abundant as photons. However, the experimental limits on the neutrino mass preclude it from
providing enough matter density to close the universe. Plenty of more exotic WIMPS have
been suggested, but as yet none has been detected. Agreat deal of experimental and theoretical
ingenuity and effort is being devoted to identifying the elusive missing matter.
The dark matter is also inextricably connected with the formation of galaxies and galaxy
clusters, since it must dominate the gravity of the universe. It is well known that galaxies tend
to occur in clusters, such as the Hercules cluster. Galaxy clusters are gravitationally bound;
that is, the galaxies within a cluster orbit one another. The Milky Way is a member of a small
cluster of a few dozen members, dominated by itself and the Andromeda galaxy; this cluster
is called the Local Group. The nearest large cluster to the Local Group is the Virgo cluster,
about 16 Mpc distant.
1.4 Problems and Mysteries
Despite its successes, the standard big bang cosmology has some problems that are difcult to
resolve. Among these are:
The horizon problem
The atness problem
The structure problem
The relic problem
These can be summarized as follows: (1) The universe is observed to be highly homogeneous
and isotropic, but how did it become so when all regions of the observable universe were not
in mutual causal contact at early times? (2) The universe is nearly at today, but this implies
that it must have had an very nearly equal to one at early times. Unless the universe is
exactly at, this seems to require ne tuning. Why is the universe so at? (3) What formed
the perturbations that lead to the structure we see around us? Why is structure the same
everywhere, even though different parts of the universe were not causally connected early in
the big bang model? (4) GUTs predict massive particles that are not observed. What happened
to these relics of the GUT epoch?
The inationary model addresses all these issues by presuming that what we call the ob-
servable universe is actually a very small portion of the initial universe that underwent a de
Sitter phase of exponential expansion around the time of the GUT epoch. This model posits
10 1 History and Overview
that what became our observable universe was small enough to be in causal contact at the
big bang; it then grew at an exponential rate during the inationary epoch. The exponential
growth had the effect of attening out any curvature, stretching the geometry of the universe
so much that it became at. Any massive GUT particles were diluted, spread out over this
now fantastically huge domain to an extremely tiny density, so that they no longer are observ-
able. Quantum uctuations in the vacuum are preserved and blown up to large scales by the
expansion, providing the seeds for structure formation.
The source of this exponential growth was a negative pressure produced by a nonzero
vacuum energy. A nonzero vacuum energy could result from quantum processes in the early
universe. In quantum eld theory, a eld is associated with each particle, and the eld in turn
is related to a potential, the latter being a function which describes the energy density of the
eld. The right potential would result in a false vacuum, a situation in which the eld was zero
but the corresponding potential was not zero. The false vacuum state could have provided a
vacuum energy that would behave exactly like a positive (repulsive) cosmological constant,
resulting in a temporary de Sitter phase during which the patch of universe grew by a factor
of perhaps 10
100
or even more. Eventually, however, this vacuum energy was converted into
real particles and the eld found its way to the true vacuum, bringing the ination to a halt.
The universe then continued to evolve from this point as in our standard model.
The inationary model is an area of active research. It makes some predictions about
the structures in the universe which are consistent with the CMBR data, although not yet
proven, and it predicts that the present universe should be at. This may present a problem
for the model since most measurements give < 1. Furthermore, = 1 models are too
young for the larger Hubble values that tend to be measured by recent techniques. Moreover,
the particle that might have provided the vacuum energy density is still unidentied, even
theoretically; it is sometimes called the inaton because its sole purpose seems to be to have
produced ination. Despite these outstanding questions, it seems difcult to understand how
the horizon problem could be explained unless something like ination occurred. Research
continues along all these lines of investigation.
1.5 The Edge of Time
Completely unknown is the beginning of time, the point in the big bang known as the Planck
epoch. During this time the universe was sufciently dense and energetic that all the funda-
mental forces, including gravity, were merged into one grand force. Quantum mechanics is
generally associated only with the world of the very small, but during the Planck epoch the
entire observable universe was tiny. Under such conditions, quantum mechanics and gravity
must merge into quantum gravity. Unfortunately, at this time physicists do not have a theory
of quantum gravity. So we can speculate on what such a theory might be like, and what it
might tell us.
The basis for quantum mechanics is the recognition that everything has a wavelike nature,
even those things we normally consider particles. By the same token, those things that we
usually consider waves (e.g., light) also have a particle nature. The evolution of quantum
systems is governed by the Schroedinger equation. However, the Schroedinger equation gives
only the evolution of the probabilities associated with a system. Interpreting what this means
is somewhat difcult, given our usual expectations regarding the nature of reality. According
to the Copenhagen interpretation of quantum mechanics, a system exists in a superposition of
1.5 The Edge of Time 11
states so long as it remains unobserved. An observation, which it must be understood refers
to any interaction that requires a variable to take a value and need not imply a conscious agent,
collapses the wavefunction. The collapse of the wave function means that the system abruptly
ceases to obey Schroedingers equation; what was previously probabilities becomes a known
quantity. The story of Schroedingers cat illustrates. A cat is locked in a box containing a
vial of poison, which will be broken by a quantum process such as the decay of a radioactive
atom. This event has some probability, which may be calculated, of occurring during any
particular time interval. While the cat is in the box, unobserved, we cannot know whether the
event has taken place or not and thus whether the cat is alive or dead. A strict application of
the Copenhagen interpretation demands that the cat be neither alive nor dead, or perhaps both
alive and dead, in some superposition of states according to the probability that the atom has
decayed; in this view, the cat becomes alive or dead only when the experimenter opens the box
to investigate. Schroedingers cat illustrates difculties with standard quantum mechanics
as applied to complex systems such as living beings. Schroedinger himself meant this thought
experiment to show that quantum mechanics, a young science at the time, did not apply to
such systems. Yet when we seek to apply quantum mechanics to cosmology, we know that it
must apply to the universe as a whole, and thus there is no such thing as classical behavior;
everything is ultimately quantum.
When we ponder the Planck era, we are led to questions about the nature of space and
time themselves. What is it that provides the arrow of time, the perception that we move
into the uncertain future and leave behind the unchangeable past. The laws of physics are
time symmetric, meaning that they work the same whether time runs forward or backward.
(A substitution of -t for t everywhere gives the same equations). The one exception is the
second law of thermodynamics which states that entropy must increase with time. This means
that a complicated system will tend to evolve toward its most probable state, which is a state
of equilibrium (and maximum disorder). If the sense of the arrow of time comes from the
second law this means that the big bang had to start in a state of low entropy (high order),
and the arrow of time results from the universal evolution from this initial state to the nal
disordered state, be it big crunch or the heat death of the ever expanding universe. We are led
to ask whether the theory of quantum gravity explains why the initial big bang was in a low
entropy state. Is quantum gravity a theory that is not time symmetric? Does the second law
of thermodynamics, an empirical relationship rst discovered by engineers in the nineteenth
century, tell us something about the most profound secrets of the universe?
Although we have no established theory of quantum gravity, some promising starts have
been made. One of the most studied is string theory, in which reality at the Planck scale
of distance and time is described by the quantum oscillations of strings and loops. String
theories require that many more spatial dimensions exist than our familiar three. At least
ten spatial dimensions exist in these theories, but only three are of cosmic scale; the rest
are compactied into coils the size of the Planck distance. Thus the very early universe
may have undergone a proto-ination in which three spatial dimensions grew into those
that make up the observable universe. String theories are not yet well understood and many
details remain to be worked out, but so far they can at least unify gravitation and the other
fundamental forces in a natural way. They are not the only candidates, however; another
theory envisions that the exotic level of the Planck scale consists of a foam of quantized space
and time. Perhaps someday we shall have a better understanding of the mysteries of the Planck
scales. Such a discovery would be at least as momentous as general relativity itself.
12 1 History and Overview
1.6 Textbooks
Cosmological Physics, by John Peacock (CUP 1998)
Principles of Physical Cosmology, by P. J. E. Peebles (Princeton University Press,
Princeton, 1993)
The Early Universe, by Kolb and Turner (Addison-Wesley, New York, 1990)
The Early Universe, by G. B orner (Springer-Verlag 2003)
Modern Cosmology, by S. Dodelson (Academic Press 2003)
An Introduction to Modern Cosmology, by A. Liddle (John Wiley 2003)
Physical Foundations of Cosmology, by V. Mukhanov (CUP Nov. 2005)
1.7 Units
Physical Units: The speed of light is c = 299 792 458 m/s, one light second is therefore
299,792 km. Velocity is usually taken in dimensionless units, i.e. in units of the speed of
light. With the Boltzmann constant k
B
= 1.3805 10
23
J/K one can convert temperature
into energy units
1 eV = 11, 600 K = 1.78 10
36
kg = 1.60 10
19
J . (1.1)
Astronomical Units: The natural unit of mass is the solar mass M

= 1.99 10
30
kg,
and the common unit of length is one parsec, 1 pc = 3.26 light years = 3.09 10
16
m. More
common in cosmology is 1 Mpc = 10
6
pc, which is a typical distance between galaxies. We
also use 1 Gpc = 1000 Mpc for the scale of the Universe.
The Hubble constant H
0
is usually given in units of 100 km/s/Mpc in the form of H
0
=
100 h km/s/Mpc. Since the Hubble constant is a velocity per distance, the inverse of the
Hubble constant denes then the Hubble time t
H
= 1/H
0
, which is equal to 13.97 billion
years for H
0
= 72. Similarly, R
H
= c/H
0
denes a scale of the Universe, which is called the
Hubble radius. R
H
= 4300 Mpc for the standard value of the Hubble constant.
2 The Observable Universe
The observable universe is the space around us bounded by the event horizon the distance to
which light can have traveled since the universe originated. This space is huge but nite with
a radius of 4200 Mpc. There are denite total numbers of everything in this volume: about
10
11
galaxies, 10
21
stars, 10
78
atoms, 10
88
photons. There is a hierarchy of structure: Every-
thing is composed of smaller things and is a part of something larger as shown in Figure 2.1.
The character of structures with different scale changes according to the interplay of various
physical forces. Quantum phenomena control the small scales, while gravity dominates on
large scales, and both come into play at the beginning of the universe.
Note that according to inationary cosmology, the entire universe is much bigger than
the observable one, and the conne of observable universe depends on the location. Observers
living in the Andromeda galaxy and beyond have their own observable universes that are
different from but overlap with ours.
In a certain sense, we live in the centre of the universe that we observe, somewhat in
contradiction to the Copernican principle, which says that the Universe is more or less uniform
and it has no distinguished centre. This is simply because light does not travel innitely fast
and we must make observations of the past. As we look further and further away, we see
things from epochs (times) closer and closer to the limit of time zero in the Big bang model.
Because light travels at the same speed in any direction towards us, we live at the geometrical
centre of our own observable universe.
In 1922, Alexander Friedmann predicted the Big Bang cosmology, which portrays the
universe as expanding space from a point where the matter-energy density was extremely high.
The expansion can be visualized by a two dimensional analogy. As the balloon expands, all the
points on the surface recede from each other, and the wavelength on the surface is stretched. It
is similar to the shift to longer wavelength when the source and receiver are moving away from
each other. This phenomenon is called red shift of the spectrum because in visible light the
shift to longer wavelength is toward the red colour. It plays a prominent role in discovering the
cosmic expansion through the detection of the spectral line shift from distant galaxies. Note
that contrary to the balloon analogy, it is the space itself that is expanding. It needs neither a
center to expand away from nor empty space on the ouside to expand into.
There are four key observational facts which require the Universe to be described in terms
of a relativistic model:
the expansion of the Universe (also called Hubbleexpansion);
cosmic microwave radiation (CMBR) which is a relic from the recombination era and,
the distribution of galaxies in the Universe which results from the gravitational action of
small perturbations rooted in the very early epoch of the Universe;
14 2 The Observable Universe
Figure 2.1: The observable Universe as seen from the Earth. This extends in a way the Aris-
totelian cosmos of Fig. 1.1. More realistically, one should center the observable Universe on the
solar system barycenter. The farther out we look, the farhter back in time we see. Light takes 50
million years to arrive from M 87. The limit of our view is the time when the Universe emerged
from a state of hot plasma and became transparent, some 300,000 years after the big bang.
the presence of dark matter whose gravity dominates the motion in galactic halos and
galaxy clusters. Though the nature of dark matter is still completely obscure, its existence
can no longer be avoided.
When Einstein developed his theory of gravity in the general theory of relativity, he
thought he ran into the same problem that Newton did: his equations said that the universe
should be either expanding or collapsing, yet he assumed that the universe was static. His
original solution contained a constant term, called the cosmological constant, which cancelled
the effects of gravity on very large scales, and led to a static universe. After Hubble discov-
ered that the universe was expanding, Einstein called the cosmological constant his greatest
blunder.
At around the same time, larger telescopes were being built that were able to accurately
2.1 The Universe is Expanding and Accelerating 15
Figure 2.2: The observable Universe of Superclusters. Superclusters are large groupings of
smaller galaxy groups and clusters, and are among the largest structures of the cosmos. The ex-
istence of superclusters indicates that the galaxies in our Universe are not uniformly distributed;
most of them are grouped together in groups and clusters, with groups containing up to 50
galaxies and clusters up to several thousand. Those groups and clusters and additional isolated
galaxies in turn form even larger structures called superclusters. The typical distance between
superclusters is about 100 Mpc. No clusters of superclusters are known, but the existence of
structures larger than superclusters is debated (Filaments).
measure the spectra, or the intensity of light as a function of wavelength, of faint objects. Us-
ing these new data, astronomers tried to understand the plethora of faint, nebulous objects they
were observing. Between 1912 and 1922, astronomer Vesto Slipher at the Lowell Observatory
in Arizona discovered that the spectra of light from many of these objects was systematically
shifted to longer wavelengths, or redshifted. A short time later, other astronomers showed that
these nebulous objects were distant galaxies.
2.1 The Universe is Expanding and Accelerating
Until 1929, the Universe of galaxies was thought to be static. Even Einstein originally searched
for a static solution of his euqations. For that reason, he had to introduce the cosmological
constant which is nowadays fundamental for the understanding of the dynamics of the Uni-
16 2 The Observable Universe
verse.
In 1929 Edwin Hubble, working at the Carnegie Observatories in Pasadena, California,
measured the redshifts of a number of distant galaxies. He also measured their relative dis-
tances by measuring the apparent brightness of a class of variable stars called Cepheids in
each galaxy. When he plotted redshift against relative distance, he found that the redshift z of
distant galaxies increased as a linear function of their distance d
cz = H
0
d . (2.1)
The only explanation for this observation is that the Universe of galaxies is expanding.
Figure 2.3: The mean apparent magnitude of Cepheids in a local galaxy as a function of the period.
[Data: Labhardt]
For the past 70 years astronomers have sought a precise measurement of Hubbles constant
H
0
, ever since astronomer Edwin Hubble realized that galaxies were rushing away from each
other at a rate proportional to their distance, i.e. the farther away, the faster the recession. For
many years, right up until the launch of the Hubble telescope the range of measured values
for the expansion rate was from 50 to 100 kilometers per second per megaparsec. Local
galaxies, i.e. galaxies between us and the Virgo and Hercules cluster, are not very suitable to
test the assumption of expansion (see Fig. 2.4). Since Cepheids can only be used as standard
candles upto distances of the Virgo galaxies, Supernovae of type Ia turned out to be much more
suitable. They become as bright as normal galaxies and are therefore visible upto cosmological
distances (Fig. 2.5). The main goal of the HST Key Project on the Extragalactic Distance
Scale was to determine the Hubble Constant, H
0
, to an accuracy of +/- 10%. This goal has
been achieved by the systematic observations of Cepheid variable stars in several galaxies
2.1 The Universe is Expanding and Accelerating 17
Figure 2.4: The recession velocities as a function of distances determined by Cepeheid distances for
local galaxies. The observed redshift of the galaxies is corrected by means of the motion against the
cosmic microwave background. [HST Key Project]
Figure 2.5: The distance module as a function of redshift z, as tested with SN Ia.
using the Hubble Space Telescope. The Key Project team currently had 28 members [5].
Cepheid distances lie at the heart of the HST Key Project on the Extragalactic Distance Scale.
The Key Project has been designed to use Cepheid variables to determine primary distances
18 2 The Observable Universe
to a representative sample of galaxies in the eld, in small groups, and in major clusters. The
galaxies were chosen so that each of the secondary distance indicators with measured high
internal precisions can be accurately calibrated in zero point, and then intercompared on an
absolute basis. The Cepheid distances can then be used for secondary calibrations and applied
to independent galaxy samples at cosmologically signicant distances. Cepheid distances to
the Virgo and Fornax clusters provide a consistency check of the secondary calibrations.
Type Ia Supernova are the explosions of white dwarfs. This is a pinnacle that only a few
stars like our sun are able to achieve. Unfortunately we are not sure exactly how these events
occur. We think they are related to white dwarf stars which are near another star in a binary
system. Chandrasekhar, as part of his Nobel Prize in Physics demonstrated that white dwarf
stars, if they become more massive than 1.4 times our sun can explode. They do this because
at this point, the forces (electrons repelling electrons) which keep the star from collapsing
against the force of gravity, lose their battle, and the white dwarf begins to collapse. As you
may recall, white dwarf stars are not made of Iron, (instead they are composed of Carbon
and Oxygen) and there is still substantial amounts of nuclear energy left in their atoms. As
the white dwarf begins to collapse against the weight of gravity, this material is ignited, and
rather than collapsing further, this nuclear blast wave consumes the star in a second, creating
an explosion 10 to 100 times brighter than a Type II supernova.
The idea to measure the Universe with Supernovae is not new, it has long been contem-
plated, but it is only in the past decade that it has become feasible. The rst distant SN Ia
was discovered in 1988 by a Danish team, but it wasnt until 1994 that they were discovered
in large numbers. Since 1995 two teams have been discovering these objects: the High-Z SN
Search, and the Supernova Cosmology Project.
For redshifts z < 0.1, curvature effects are not important and the Hubble relation cz =
H
0
d can be used in its original version for the distance module (Fig. 2.5)
mM = 5 log(d/pc) 5 = 5 log(z) 5 + 5 log(c/H
0
) . (2.2)
SN Ia are observable to medium redshifts z < 1.5 giving Hubble diagrams which demonstrate
that the classical Hubble relation fails at least beyound redshift 0.2 (Fig. 2.6). The deviations
from the straight line are the effects of the expansion of the Universe. They can only be
understood within a relativistic cosmological model.
The Universe is Accelerating
Recent studies of Type Ia supernovas, including measurements by the Supernova Cosmology
Group led by Saul Perlmutter at LBNL and the High Z-SN Search team lead by Brian P.
Schmidt, have produced signicant evidence that over cosmological distances they appear
dimmer than would be expected if the universes rate of expansion was constant or slowing
down. This was the rst direct experimental evidence for an accelerating universe potentially
driven by a positive Cosmological Constant. However, only about 80 supernovas accumulated
over several years have been studied and other explanations have not been completely ruled
out.
A space mission is now being considered that would increase the discovery rate for such
supernovas to about 2000 per year. Discovery of so many more supernova would help elimi-
nate possible alternative explanations, give experimental measurements of several other cos-
mological parameters, and put strong constraints on possible cosmological models. The satel-
lite called SNAP (Supernova / Acceleration Probe) would be a space based telescope with
2.1 The Universe is Expanding and Accelerating 19
Figure 2.6: The redshift of SN Ia as a function of the luminosity distance derived from the distance
module for a set of recent data. At redshifts z > 0.15, the Hubble relation fails. The lines correspond
to luminosity distances of various cosmological models which will be discussed later on. [Data: Tonry
2003]
a one square degree eld of view with 1 billion pixels. Such a satellite would also comple-
ment the results of proposed experiments to improve measurements of the cosmic microwave
background.
In addition to the supernova discover programitself, LBNLs Supernova Cosmology Group
has unique expertise in large charge-coupled (CCD) detectors. While smaller CCDs are now
common, LBNL has developed techniques to construct the large mosaics required for SNAP
by stitching together several hundred of the largest ones. The group has also devised a way
to manufacture the detectors at signicantly reduced cost. Technically, the CCDs have high
resistivity with excellent quantum efciency at 1 micron, which is the same as the emission
from distant Type Ia supernova and where conventional CCDs have very low sensitivity.
The SNAP baseline science objective is to obtain a high statistics calibrated dataset of Type
Ia supernovae to redshifts of 1.7 with excellent control over systematic errors. The statistical
sample is to be 2 orders of magnitude greater than the current published set of 42 supernovae,
and is to extend much farther in distance and time. From this dataset we expect to obtain
a 2% measurement of the mass density of the universe, a 5% measurement of the vacuum
energy density, a 5% measurement of the curvature, and a 5% measurement of the equation
of state of the dark energy driving the acceleration of the universe. Systematic studies
will include a measurement of the reddening of spectra from ordinary dust at redshifts
up to 1.7, and studying potential grey dust sources. Using type Ia supernovae as standard
candles will require measurement of the key luminosity indicators: the light curve peak and
width. The redshift of the host galaxy of the supernova needs to be measured, supernova type
20 2 The Observable Universe
Figure 2.7: The Supernova/Acceleration Probe (SNAP) satellite observatory is capable of mea-
suring thousands of distant supernovae and mapping hundreds to thousands of square degrees of
the sky for gravitational lensing each year. The results will include a detailed expansion history
of the universe over the last 10 billion years, determination of its spatial curvature to provide
a fundamental test of ination, precise measures of the amounts of the key constituents of the
universe,
M
and

, and the behavior of the dark energy and its evolution over time. [SNAP
Homepage]
idented, and spectral features studied. Effects correlated with host galaxy morphology and
the position of the supernova in the host galaxy will also be studied. These properties may
indicate diferences in stellar population from which the supernova came and therefore can
be used to test whether the intrinsic brightness of the supernova changes systematically with
redshift.
The Supernovae Legacy Survey [1] is comprised of two components: an imaging survey
to detect SNe and monitor their light-curves, and a spectroscopic program to conrm the
nature of the candidates and to measure the redshift. The imaging is taken as part of the deep
CFHT Legacy Survey using the one sqare degree imager MegaCam. This program has been
allocated 474 nights over 5 years. Followup spectroscopy uses various 810 meter class
telescopes.
Distance measurements to 71 high redshift type Ia supernovae discovered during the rst
year of the 5-year Supernova Legacy Survey (SNLS) are presented in [1]. These events were
detected and their multi-color light-curves measured using the MegaPrime/MegaCam instru-
ment at the Canada-France-Hawaii Telescope (CFHT), by repeatedly imaging four one-square
degree elds in four bands. Follow-up spectroscopy was performed at the VLT, Gemini and
Keck telescopes to conrm the nature of the supernovae and to measure their redshift. With
this data set, they have built a Hubble diagram extending to z = 1 (shown in Fig. 2.8), with all
2.1 The Universe is Expanding and Accelerating 21
SN Redshift
0.2 0.4 0.6 0.8 1
B

34
36
38
40
42
44
)=(0.26,0.74)

,
m
(
)=(1.00,0.00)

,
m
(
S
N
L
S
1
s
t Y
e
a
r
SN Redshift
0.2 0.4 0.6 0.8 1

)
0

H
-
1

c
L

(

d
1
0

-

5

l
o
g
B

-1
-0.5
0
0.5
1
Figure 2.8: Hubble diagram of SNe from the SN Legacy program [1]. The bottom plot shows
the residuals for the best t to a at Lambda cosmology.
distance measurements involving at least two bands. Essentially,
B
= m
B
M is the dis-
tance modulus for an absolute magnitude of SNIa M = 19.31 0.03 +5 log h
70
, corrected
for stretching and color. Systematic uncertainties are evaluated making use of the multi-band
photometry obtained at CFHT. Cosmological ts to this rst year SNLS Hubble diagram give
the following results:

M
= 0.2630.042 (stat) 0.032 (sys) for a at LambdaCDM model; and w = 1.023
0.090 (stat) 0.054 (sys) for a at cosmology with constant equation of state w, when com-
bined with the constraint from the recent Sloan Digital Sky Survey measurement of baryon
acoustic oscillations. As a result, the Universe is denitely accelerating, and not decelerat-
ing.
22 2 The Observable Universe
2.2 Cosmic Background Radiation as a Relic of the Early
Universe
In 1963, Arno Penzias and Robert Wilson, two scientists in Holmdale, New Jersey, were work-
ing on a satellite designed to measure microwaves. When they tested the satellites antenna,
they found mysterious microwaves coming equally from all directions. At rst, they thought
something was wrong with the antenna. But after checking and rechecking, they realized that
they had discovered something real [11]. What they discovered was the radiation predicted
years earlier by Gamow [6], Alpher and Herman [1]. The radiation that Penzias and Wilson
discovered, called the Cosmic Microwave Background Radiation (CMBR), convinced most
astronomers that the big bang theory was correct. For discovering the Cosmic Microwave
Background Radiation, Penzias and Wilson were awarded the 1978 Nobel Prize in Physics.
The signicance of the discovery of the CMBR is that it was predicted by the basic phys-
ical theory of the expanding universe, rst laid down decades before, by Georges Lematre.
While the cosmology wars of the 1940s and 1950s pitted several competing models against
one another, only the Big Bang had predcited the necessity of the CMBR as a consequence of
the theory. So its only natural that its discovery gave support to that theory. The supporters
of alternative cosmologies tried to bring such a background radiation out of their models after
the fact, but always with unsatisfying, arbitrary tricks. Only the Big Bang actually requires
that there be a CMBR. The existence of the CMBR was one of several strong reasons that
developed during the 1960s & 1970s, which drove Big Bang cosmology into its premiere
position amongst scientists.
Subsequent observations conrmed, as well as it could be conrmed by limited, ground
based observations, that the CMBR discovered by Penzias & Wilson indeed had the required
thermal spectrum, and was indeed isotropic. These two facts alone are strong indicators that
the CMBR is the predicted relic of Lematres primeval atom, and argues against alternative
cosmologies.
After Penzias and Wilson found the Cosmic Microwave Background Radiation, astro-
physicists began to study whether they could use its properties to study what the universe was
like long ago. According to Big Bang theory, the radiation contained information on how
matter was distributed over ten billion years ago, when the universe was only 500,000 years
old. At that time, stars and galaxies had not yet formed. The Universe consisted of a hot soup
of electrons and atomic nuclei. These particles constantly collided with the photons that made
up the background radiation, which then had a temperature of about 3000 K.
Soon after, the Universe expanded enough, and thus the background radiation cooled
enough, so that the electrons could combine with the nuclei to form atoms. Because atoms
were electrically neutral, the photons of the background radiation no longer collided with
them.
When the rst atoms formed, the universe had slight variations in density, which grew into
the density variations we see today - galaxies and clusters. These density variations should
have led to slight variations in the temperature of the background radiation, and these varia-
tions should still be detectable today. Scientists realized that they had an exciting possibility:
by measuring the temperature variations of the Cosmic Microwave Background Radiation
over different regions of the sky, they would have a direct measurement of the density vari-
ations in the early universe, over 10 billion years ago. In 1990, the a satellite called the
Cosmic Microwave Background Explorer (COBE) measured background radiation tempera-
tures over the whole sky (Fig. 2.9). COBE found variations that amounted to only about 5
2.2 Cosmic Background Radiation as a Relic of the Early Universe 23
Figure 2.9: The spectrum measured by COBE is a perfect Planckian distribution.
Figure 2.10: The motion of the Earth generates a dipole anisotropy T/T 0.001 in the temperature
distribution on the Sky as observed by COBE.
parts in 100,000, but revealed the density uctuations in the early universe (Fig. 2.10).
The initial density variations would be the seeds of structure that would grow over time to
become the galaxies, clusters of galaxies, and superclusters of galaxies observed today by the
Sloan Digital Sky Survey. With the Sloan data, along with data from COBE, astronomers will
be able to reconstruct the evolution of structure in the universe over the last 10 to 15 billion
years. With this information, we will have a deep understanding of the history of the universe,
24 2 The Observable Universe
Figure 2.11: Temperature uctuations on the Sky as observed by WMAP. The typical angular separation
of temperature spots is about one degree, the resolution of the instrument is 13 arcminutes, in contrast to
the 7 degree resolution for the COBE instrument. [Source: Tegmark]
Figure 2.12: Harmonic analysis on the sphere: Dipole anisotropy (l = 2) and anisotropy mit
l = 16 demonstrate, how a complicated uctuation spectrum can be decomposed into a super-
position of all possible lvalues.
which will be an almost unbelievable scientic and intellectual achievement.
The cosmic microwave radiation is emitted by a sphere (the socalled Last Scattering Sur-
face at redshift z = 1070). We can therefore consider the temperature as a scalar function on
the sphere, which can be decomposed into spherical harmonics [9]
T(, ) = T(, ) T
0
=
l
max

l=0
m=+l

m=l
a
lm
Y
lm
(, ) (2.3)
Y
lm
are the spherical harmonics (as used e.g. in Quantum mechanics). l
max
follows from the
resolution on the sky, which is 7 degrees for COBE, 10 arcminutes for MAXIMA1, and 13
arcminutes for WMAP. The functions Y
lm
form a complete basis on the sphere, i.e.
_
4
Y

lm
(, ) Y
LM
(, ) d =
lL

mM
. (2.4)
2.2 Cosmic Background Radiation as a Relic of the Early Universe 25
Figure 2.13: The power spectrum of the CMBR temperature uctuations as a function of mul-
tipoles l, i.e. as a function of angular separation before WMAP. The angular scale is given at
the upper axis of the gure. COBE had an angular resolution of 7 degrees, ballon experiments
such as Boomerang and Maxima could determine uctuations upto l 800, though with con-
siderable error bars due to the limited sky coverage. The solid line gives the impression of
uctuations expected in a CDM model. Here, the greatest uctuations are expected on scales of
about one degree. On much smaller scales, uctuations decay very rapidly.
From this orthogonality relation we obtain the representation of the coefcients a by means
of N observed values, i.e. the temperature in N discrete directions on the sphere. The sample
function values T
p
can then be used to estimate the parameters a
lm
a
lm
=
N1

p=0
T(
p
,
p
) Y

lm
(
p
,
p
) . (2.5)
26 2 The Observable Universe
The zeroeth of the spherical harmonics Y
lm
divide up the sphere into invidual cells, with
an extension near the equatorial plane of = /l. This is the meaning of the multipole
Figure 2.14: The power spectrum measured by WMAP. This gure summarizes the experi-
mental results as of March 2003. The existence of the resonance peak at l 200 has been
conrmed, the secondary maximum is barely visible. [Data: Tegmark, WMAP]
of order l. The power represented by a multipole of order l is obtained by averaging over all
mvalues
C
l

1
2l + 1
m=+l

m=l
a
lm
a

lm
. (2.6)
The parameters C
l
represent indeed a complete statistical representation of Gaussian temper-
ature uctuations. The value of l = 200 just corresponds to an angular separation of about
one degree on the sky. COBE could only resolve scales upto l = 20, balloon measurements
had a bigger resolution, but they could not resolve structure beyond l = 750 (Fig. 2.14).
WMAP (Wilkinson Microwave Anisotropy Probe) now has conrmed these earlier mea-
surements with an unprecented accuracy and a resolution of 13 arcminutes [3] (Fig. 2.14).
2.2 Cosmic Background Radiation as a Relic of the Early Universe 27
Figure 2.15: An image of the intensity and polarization of the cosmic microwave background
radiation made with the Degree Angular Scale Interferometer (DASI) telescope. The small tem-
perature variations of the cosmic microwave background are shown in false color, with yellow
hot and red cold. The polarization at each spot in the image is shown by a black line. The
length of the line shows the strength of the polarization and the orientation of the line indicates
the direction in which the radiation is polarized. The size of the white spot in the lower left
corner approximates the angular resolution of the DASI polarization observations. [Data: DASI
collaboration]
The skymap data products derived from the WMAP observations have 45 times the sensitiv-
ity and 33 time the angular resolution of the COBE DMR mission. In particular, the peak
at l = 200 in the power spectrum is extremely important for the reduction of cosmological
parameters. But in addition, the secondary peak is also visible, however with some bigger
uncertainty. The data brings into high resolution the seeds that generated the cosmic structure
we see today. These patterns are tiny temperature differences within an extraordinarily evenly
dispersed microwave light bathing the Universe, which now averages a frigid 2.73 degrees
above absolute zero temperature. WMAP resolves slight temperature uctuations, which vary
by only millionths of a degree [8].
Polarisation of the CMBR: The polarization of the cosmic microwave background (Fig.
2.15) was produced by the scattering of cosmic light when it last interacted with matter, nearly
14 billion years ago. If no polarization had been found, astrophysicists would have to reject
28 2 The Observable Universe
all their interpretations of the remarkable data they have compiled in recent years.
Figure 2.16: Triangulation of spherical surfaces with the HealPix procedure. The surface of a
sphere is primarily divided up into 12 domains of equal area. These segments are then divided
up in ahierarchical procedure. top right the topography of the Earth is shown with 3,145,728
pixels (corresponding to a resolution of 7 arc minutes) and left top the anisotropy of the CMBR
is given for 12,582,912 pixels (resolution of 3.4 arc minutes).
Triangulation of Spherical Surfaces: Measurements of the temperature T(n) in a given
direction n on the Sky means that we have to treat a scalar function on the sphere (Fig. 2.16).
This requires an efcient triangulation of the 2sphere in the sense that all pixels have about
the same area. A simple discretisation of the angles and produces pixel areas which
2.2 Cosmic Background Radiation as a Relic of the Early Universe 29
become smaller and smaller towards the polar regions.
Numerically efcient procedures should have the following properties:
Equal area for all pixels.
Pixels are distributed on lines of constant latitude. This property is essential for all har-
monic analysis applications involving spherical harmonics. Due to the iso-latitude distri-
bution of sampling points the speed of computation of integrals over individual spherical
harmonics scales as

N with the total number of pixels, as opposed to the Nscaling


for the non-iso-latitude sampling distributions (examples of which are the Quadrilateril-
ized Spherical Cube used for the NASAs COBE data, and any distribution based on the
symmetries of the icosahedron [13]).
The sphere is hierarchically tessellated into curvilinear quadrilaterals. The lowest resolu-
tion partition is comprised of 12 base pixels. Resolution of the tessellation increases by
division of each pixel into four new ones. The Figure 2.16 illustrates (clock-wise from
upper-left to bottom-left) the resolution increase by three steps from the base level, i.e.
the sphere is partitioned, respectively, into 12, 48, 192, and 768 pixels.
The HealPix algorithm (Hierarchical Equal Area isoLatitude Pixelisation) satises these re-
quirements (Fig. 2.16) [7]. Each healpix sphere contains 12 N
2
pixel. Typical values are
N = 256, 512, 1024. This triangulation of the spherical surface has many application in as-
trophysics, in particular it is very useful for the numerical solution of the radiative transfer
equation.
Future CMBR Missions: Planck (formerly COBRAS/SAMBA) was ofcially adopted as
the third Medium-Sized Mission of ESAs Horizon 2000 scientic program in November 1996
with a scheduled launch in late 2005, now postponed to 2007. Planck currently consists of 4
frequency channels (30-100 GHz) that use HEMT ampliers running at 20K and 6 channels
(100-857 GHz) that use bolometer detectors in a 0.1 K dewar. The single 1.3m x 1.5m primary
reector produces a 0.1 degree beam at 200 GHz and above. The sensitivity of PLANCK is
almost an order of magnitude higher than WMAP.
Summary
WMAP provided higher accuracy measurements of many cosmological parameters than
had been available from previous instruments.
The Hubble constant is 71 4 km/s/Mpc.
The universe is composed of
1. 4% ordinary matter,
2. 23% of an unknown type of dark matter (DM),
3. and 73% of a mysterious dark energy (DE).
4. The Age of the Universe is (13.7 0.2) billion years.
This is a conrmation of the so-called concordance Lambda-CDM model.
Besides this, WMAP measured many more parameters, which will be discussed later on.
30 2 The Observable Universe
2.3 Galaxies are not Uniformly Distributed
Matter in the universe is not distributed randomly. Galaxies, quasars, and intergalactic gas
outline a pattern that has been compared to soap bubbles - large voids surrounded by thin
walls of galaxies, with dense galactic clusters where walls intersect. One of the primary goals
Figure 2.17: The galaxy distribution in the CfA redshift survey. Each dot represents a galaxy. Redshift
runs along the radius of the circle, we are located in the center of the circle. The radius of the circle
amounts to 15,000 km/s, the Coma cluster and the Great Wall are at about 8,000 km/s. The center is part
of the Virgo cluster.
of modern surveys (CfA, Las Campanas, 2dF, SDSS, 6dF) is to map this structure in great
detail, out to large distances. There are many theories about how the universe evolved, and
the theories predict different large-scale structrues for the universe. The SDSSs map may tell
us which theories are right - or whether we will have to come up with entirely new ideas.
Galaxies are usually found near each other, in galactic clusters. The distribution of these clus-
ters, and how this distribution evolves with time, are important tests of cosmological models:
for instance, different cosmological models predict different numbers of galaxy clusters at dif-
ferent redshifts. Additionally, not only are galaxies clustered, but the clusters themselves are
2.3 Galaxies are not Uniformly Distributed 31
Figure 2.18: The galaxy distribution in the 2dF galaxy survey for a strip of 3 degrees wide in declination.
Galaxies are detected out to redshifts of 0.2, roughly corresponding to a distance of one Gigaparsec.
Figure 2.19: The redshift distribution in the 2dF galaxy redshift survey. Beyond redshift 0.15, only
the most luminous galaxies can be detected. The solid line is the theoretical expectation based on the
Schechter form for the galaxy luminosity function.
clustered! The degree to which both galaxies and clusters tend to group together is also a test
of different theories. By studying the masses, distributions, and evolution of galaxy clusters,
32 2 The Observable Universe
we can learn something about the formation of mass in the universe: a fundamental goal of
cosmology.
Superclusters are simply clusters of galaxy clusters. Whereas clusters are typically found
in the laments and walls of the universes soap bubble, superclusters are at the intersections
of the walls. Superclusters are the largest known structures in the universe, with some as
large as 200,000,000 light-years! However, because these structures are very rare, only a
few are known. The most famous superclusters are nearby, including the Great Wall and the
Perseus-Pisces supercluster. There has been recent evidence for superclusters at redshifts of
about 1, which places important constraints on structure formation and cosmological models.
Additionally, the M/L ratios of superclusters are similar to those of clusters. This discovery
implies that the mysterious dark matter cannot contribute more to the mass of the universe
than it contributes to the mass of clusters.
Figure 2.20: A comparison between CfA and SDSS galaxy surveys. A kind of great wall is now visible
in the SDSS map, however at much larger scales.
Future Surveys: The VIRMOS-VLT Deep Survey (VVDS) is a breakthrough spectroscopic
survey which will provide a complete picture of galaxy and structure formation over a very
2.4 Gravity is Dominated by Dark Matter 33
broad redshift range (0 < z < 5) over sixteen square degrees of the sky in four separate
elds. This ambitious survey is possible thanks to the impressive multiplex gains of the VIR-
MOS instruments (VIMOS and NIRMOS) built by the Franco-Italian VIRMOS consortium
for ESO-VLT. The survey will be carried out over the consortiums 120 nights of guaranteed
time, and will contain in total a sample of 150,000 redshifts. This unique database will enable
us to trace back the evolution of galaxies, active galactic nuclei and clusters to epochs where
the universe was a fraction (about 20 per cent) of its current age. The VVDS will be compa-
rable in size to the largest redshift surveys currently underway, but will probe to much higher
redshifts. It will provide an unparalleled description of how structures and galaxy populations
evolved in the universe from high redshift to the present day.
The distribution of galaxies at higher redshifts (z 1) is the topic of deep surveys (HST
deep elds, FORS deep elds, Subaru deep elds etc). The spatial distribution at these dis-
tances is the topic of the next decade in Astronomy.
Summary
The observations (d2F and SDSS) demonstrate for the rst time that the Universe of
galaxies is homogeneous on scales beyond 100 Mpc, but denitely not within 100 Mpc.
Galaxies are assembled into groups, clusters and superclusters.
The space between clusters is mainly empty of galaxies. These regions are called voids.
2.4 Gravity is Dominated by Dark Matter
The nature and identity of the dark matter in the Universe is one of the most challenging
problems facing modern cosmology. The problem has been for the rst time formulated by
Zwicky. Given the distribution of galaxies with total luminosity L, (L) dL, one can compute
the mean luminosity density of galaxies
L =
_
(L) dL (2.7)
which is then determined by
L (2 0.2) 10
8
h
0
L

Mpc
3
. (2.8)
In the absence of a cosmological constant, one can dene a critical mass density

c
3H
2
0
/8G = 1.88 10
29
h
2
g cm
3
. (2.9)
With this, we can dene a critical masstolight ratio
(M/L)
c
=
c
/L 1390 h(M

/L

) . (2.10)
For the standard value H
0
= 70 we nd then (M/L)
c
1000 (M

/L

). With this we obtain


a value for the density parameter

m
=

c
=
(M/L)
(M/L)
c
. (2.11)
34 2 The Observable Universe
It turns out that masstolight ratios are strongly dependent on distance scale over which they
are determined: In the solar neighborhood M/L 1, yielding values of
m
0.001. In
the central parts of galaxies one nds M/L (10 20)h so that
m
0.01 for galaxies.
On larger scales (i.e. for binaries and small groups of galaxies), M/L (60 180)h with

m
0.1. nally, on the scale of galaxy clusters, M/L still increases and may be as large as
(200500)h giving
m
0.30.4. Thus on the scales of clusters, dark matter is inevitable.
Dark Matter in Clusters: Because galaxy clusters can be very massive (up to 10
14
times
the mass of the Sun), their gravity is strong enough to hold on to extremely hot gas, with
temperatures of millions of degrees. This gas emits radiation at X-ray wavelengths, which
can be observed by X-ray satellites like ROSAT, Chandra, and XMM. These satellites have
shown that a large fraction of clusters have structure and complicated internal motions, which
indicate that they are still evolving. Also, satellite observations have shown that the X-ray
emitting gas comprises the largest fraction of the visible mass in clusters, greater than the
sum of all the galaxies. This is a very interesting result - remember that galaxy clusters were
discovered as overdensities of galaxies, and now we know that galaxies are but a small part
of the total mass of clusters. Some astronomers have even suggested clusters without galaxies
may also exist just huge clumps of gas.
Figure 2.21: The Xray gas in the cluster Abell 2104 at redshift 0.15. The blue dots are Xray photons
detected by Chandra, the red dots optical galaxies in the cluster. The extension of the Xray emitting
gas is about 200 kpc. A few of the galaxies in the cluster are also emitting Xrays (these are Quasars).
[Source: Chandra Homepage]
2.4 Gravity is Dominated by Dark Matter 35
The cluster Abell 2104 is one of the lowest redshift clusters (z = 0.153) known to have a
gravitational lensing arc. Detailed analysis of the cluster properties such as the gravitational
potential using the X-ray data from ROSAT (HRI) and ASCA, as well as optical imaging
and spectroscopic data from the CFHT has been given. The cluster is highly luminous in the
X-ray with a bolometric luminosity of L
X
3 10
45
ergs/s and a high gas temperature of
keV. The X-ray emission extending out to at least a radius of 1.46 Mpc, displays signicant
substructure. The total mass deduced from the X-ray data under the assumption of hydrostatic
equilibrium and isothermal gas, is found to be M(r < 1.46 Mpc) 8 10
14
M

. The gas
fraction within a radius of 1.46 Mpc is 510%. The cluster galaxy velocity distribution has a
dispersion of (1200 200) km/s with no obvious evidence for substructure.
Direct evidence for dark matter in clusters follows from the observation of the distribution
of the hot baryonic gas. The hot gas in a cluster is held in the cluster primarily by the gravity
of the dark matter, so the distribution of the hot gas is determined by that of the dark matter.
By precisely measuring the distribution of X-rays from the hot gas, the astronomers were able
to make the best measurement yet of the distribution of dark matter in the inner region of a
galaxy cluster. Under the assumption of hydrostatic equilibrium and spherical symmetry, the
cluster total mass is directly related to the intracluster gas properties as
M
tot
(r) =
rk
B
T
g
(r)
m
p
G
_
d ln n
e
d ln r
+
d ln T
g
d ln r
_
. (2.12)
The gas density usually follows a standard law
n
e
(r) = n
0
_
1 + (r/r
0
)
2
_
3/2
. (2.13)
Hence the gravitational potential for a constant gas temperature is given by
(r)
0
=
3
2
0
2
ln[1 + (r/r
0
)
2
] (2.14)
where
2
0
= k
B
T
g
/m
p
and the total mass is given by
M
tot
(r) =
3
2
0
r
0
G
(r/r
0
)
3
1 + (r/r
0
)
2
. (2.15)
If we consider galaxies as test particles in the cluster potential well, then Jeans equation
for a collisionless, steady state, non-rotating spherically symmetric system gives
M
tot
(r) =
r
2
r
(r)
G
_
d ln n
gal
d ln r
+
d ln
2
r
d ln r
+ 2
t
_
. (2.16)

r
is the 1D velocity dispersion of the galaxies in the cluster. By measuring the density prole
and the prole of the velocity dispersion, a direct mass determination can be obtained.
Gravity can bend light, allowing huge clusters of galaxies to act as telescopes. Almost all
of the bright objects in the above Hubble Space Telescope image are galaxies in the cluster
known as Abell 2218 (Fig. 2.23). The cluster is so massive and so compact that its gravity
bends and focuses the light from galaxies that lie behind it. As a result, multiple images of
these background galaxies are distorted into long faint arcs a simple lensing effect analogous
to viewing distant street lamps through a glass of wine. The cluster of galaxies Abell 2218 is
itself about three billion light-years away in the northern constellation Draco.
36 2 The Observable Universe
Figure 2.22: Mass proles derived from Xray observations in the Hydra cluster on the scale from a
few kpc to 200 kpc. [Source: Chandra]
Dark Matter in Galaxies: Dark matter is also required to explain the at rotation curves
of spiral galaxies. The halo of spiral galaxies must be populated by a dark component with a
density distribution of
H
1/r
2
, providing a mass distribution from the Jeans equation
M(r) V
2
r/G. (2.17)
Matter distribution in the disk cannot explain the observed rotation curves.
Dark Matter Candidates: The nature of the dark matter predicted by ination is a profound
and unresolved puzzle. We have two choices. Either the dark matter consists of ordinary,
baryonic matter, or else it consists of some more exotic form of matter. The history of the
universe during the rst few minutes provides an interesting measure of the total amount of
baryonic matter in the universe that may help resolve the puzzle.
For a signicant clue to the composition of the dark matter, we look to the abundance of
the heavier isotope of hydrogen, weighing twice the mass, called deuterium, created during
the big bang. There is no alternative source for the extra deuterium other than the big bang,
since stars destroy deuterium rather than produce it. By now, a considerable fraction of any
primordial deuterium present at the birth of the galaxy would have been destroyed inside stars.
This is conrmed by observation: interstellar clouds contain deuterium, as do gravitationally-
powered stars that have not yet developed nuclear burning cores; on the other hand, evolved
stars have no deuterium.
2.4 Gravity is Dominated by Dark Matter 37
Figure 2.23: HST image of Abell 2218 with luminous arcs. The dark matter in the cluster operates as a
gravitational lense. [Data: HST]
To estimate how much deuterium was created in the big bang, one has to factor in all
the deuterium that has since been destroyed. The percentage of the isotope destroyed since
the big bang can be calculated if one knows the its rate of destruction, which can be found
by comparing the abundance of deuterated molecules in the atmosphere of Jupiter with the
abundance of deuterium in interstellar clouds. One has to choose a value for the density of
baryons that cannot exceed about a tenth of the critical density for closure of the universe,
or too little primordial deuterium would have been synthesized. Conversely, the density of
baryons cannot be too low, below 2 or 3 percent of the critical density, or else one would
overproduce deuterium, compared to what is observed in the solar system. If the universe is
at critical density, 90 percent of the matter in the universe must be nonbaryonic.
If, in a universe at critical density, most dark matter could not be baryonic, what other
forms could it take? Likely relics of the early universe are species of stable, weakly interacting
particles. One example is the neutrino, if it possesses a small mass. Normally, the neutrino
is assumed to be practically massless, but a nite mass is not implausible. There are so many
neutrinos left over from the big bang that a neutrino mass of even 50 eV, or one ten-thousandth
the mass of an electron, would sufce to close the universe. Laboratory experiments are
underway in several countries to determine a denitive mass for the neutrino, but at present
these experiments are inconclusive. The current upper limit on the electron neutrino mass,
which is obtained from tritium decay experiments, is about 1 eV. Other species of neutrinos
could have higher masses.
If the particles are very massive, possessing more mass than, say, a proton, a special name
has been coined: the WIMP, for weakly interacting, massive particle. Exotic WIMPs such
as the photino and neutralino have been postulated to exist in sufcient quantity to close the
universe. The problem is that there is no guarantee that these particles do exist. Disregarding
38 2 The Observable Universe
this uncertainty, the big bang theory predicts their density today, if they do exist and are stable
over the age of the universe.
The existence of the photino is predicted in a theory called supersymmetry. This theory
doubles the number of known particles by postulating the existence of partner -ino particles.
These particles are almost all short-lived, and exist in large numbers only in the very early
universe, when the temperature was high enough to exceed the energy scale characteristic of
supersymmetry, affectionately abbreviated to SUSY. As the universe cools, supersymmetry
is broken. The relevant energy scale is not known from theory, but it must exceed 100 GeV
to avoid conict with particle experiments. In our low-energy universe today, the lightest
supersymmetric particle should still survive. It is expected to be the partner, in the sense of
having a complementary spin, of the photon, and is therefore known as a photino. Its mass is
expected to be 10 to 100 times that of the proton. The photino is uncharged and interacts very
weakly with matter.
There is strong evidence for SUSY from experiments at CERN that measure the strength
of the nuclear interactions, which increase with increasing energy. There is no guarantee,
however, that the weak and strong nuclear force strengths will all converge to the same energy.
That they do converge at very high energy is the thesis of grand unication of the fundamental
forces, whose breakdown in the very early universe gave rise to ination. While this energy,
some 10 to the 15 GeV, is very much higher than is directly accessible by experiment, the trend
towards convergence of the disparate forces is already apparent. Only if SUSY describes the
high energy world do these three fundamental forces become indistinguishable at a unique
energy. Only therefore with SUSY could one construct a strong case for the inevitability of
grand unication.
The most natural form for dark matter is matter that we know exists, namely baryons.
The big bang explanation of the light element abundances requires the existence of baryonic
dark matter. Although these same abundances imply that most dark matter is nonbaryonic,
the amount of dark baryonic matter is still most likely several times that in luminous baryonic
matter, or about 3 percent of the critical density for closing the universe. But where do we
look for the baryonic dark matter? Ones rst expectation might be that baryonic dark matter
consists of burnt-out stars in the galactic halo, yet other forms, such as planets and black holes,
are also possible. Baryonic dark matter does exist: it is far more uncertain whether there exists
enough to solve any of the dark matter problems, that is to say, dark matter in galaxy halos,
dark matter in galaxy clusters and superclusters, or dark matter in an amount sucient to close
the universe. It is most unlikely that baryonic dark matter can account for the closure density,
as we will now see: for this, one must appeal to WIMPs, or some other weakly interacting
particle. However, baryonic dark matter is a serious candidate for dark matter at least in galaxy
halos, if not on larger scales. In acknowledgment of the rivalry between these two forms of
dark matter, the favored baryonic dark matter candidates have been dubbed MACHOs, for
massive compact halo objects.
Among the possible astrophysical objects contained in the halo are the relics of stars, dim
stars such as white dwarfs, neutron stars, or even black holes, as well as objects that have
never quite fullled themselves as stars because of their low mass. Because these objects are
invisible, or almost so, they are excellent candidates for dark matter. Moreover, MACHOs are
more natural candidates for the halo dark matter than WIMPs, because they are already known
to exist.
Two experiments reported in 1993 have found strong evidence for the existence of MA-
CHOs. The technique used is gravitational microlensing. If a MACHO in our galaxys halo
2.4 Gravity is Dominated by Dark Matter 39
passes very close to the line of sight from earth to a distant star, the gravity of the otherwise
invisible MACHO acts as a lens that bends the starlight. The star splits into multiple images
that are separated by a milliarc-second, far too small to observe from the ground. However,
the background star temporarily brightens as the MACHO moves across the line of sight in the
course of its orbit around the Milky Way halo. To overcome the low probability of observing
a microlensing event, the experiments were designed to monitor several million stars in the
Large Magellanic Cloud. Each star was observed hundreds of times over the course of a year.
A preliminary analysis of the data, taken with both red and blue lters, revealed several events
that displayed the characteristic microlensing signatures. The event durations were between
30 and 50 days.
The duration of the microlensing event directly measures the mass of the MACHO, al-
though there is some uncertainty because of the unknown transverse velocity of the MACHO
across the line of sight. The event duration is simply the time for the MACHO to cross the
effective size of the gravitational lens, known as the Einstein ring radius. The radius of the
Einstein ring is approximately equal to the geometric mean of the Schwarzschild radius of the
MACHO and the distance to the MACHO. For a MACHO half-way to the Large Magellanic
Cloud, that distance is 55 kiloparsecs. The Einstein ring radius is about equal to 1 astronomi-
cal unit, or the earth-sun distance. In order to be lensed, the MACHOs must be objects that are
smaller than the lens, so they must be smaller than an astronomical unit, roughly the radius of
a red giant star. The events detected are, to within a factor of a few, as the MACHO model of
dark halo matter predicts, and the event durations suggest a typical mass of around 0.1 solar
masses; however, there is at least a factor of 3 uncertainty in either direction.
The microlensing experiments have given robust and strong limits on the baryonic content
of the halo. Much more data from the LMC and SMC will be available soon, so we expect
the statistics to improve in the near future. The LMC events, if interpreted as due to halo mi-
crolensing, allowa measurement of the baryonic contribution to the halo, which is around 20%
for a standard halo. In this case, the most likely Macho contribution to the Milky Way halo
mass is about 8 10
10
M

, which is roughly the same as the disk contribution to the Milky


Way mass. However, the whole story has been made more complicated (and exciting) by the
much larger than expected number of bulge microlensing events. These events imply a new
component of the Galaxy, and until the nature of this new component is known, unambiguous
conclusions concerning the LMC events will not be possible. For example, if the Milky Way
disk is much larger than usually considered, a much smaller total halo mass will be required,
and so even an all-Macho halo might be allowed. Alternatively, the new Galactic component
which is giving rise to the bulge events, may also be giving rise to the LMC events, and the
Macho content of the halo could be zero. Fortunately, much more data is forthcoming, and
many new ideas have been proposed. Microlensing is fast becoming a new probe of Galactic
structure, and, beside the original potential to discover or limit dark matter, may well produce
discoveries such as extrasolar planetary systems.
Summary
The existence of Dark Matter (DM) is unavoidable in modern cosmology.
Dark Matter cannot exist of dark baryons.
There is no favorable candidate for Dark Matter.
Dark Matter and Dark Energy are the biggest mysteries for modern cosmology.
40 2 The Observable Universe
2.5 Exercises
Hubble constant: Besides Cepheids and Supernovae, many other methods are used to esti-
mate the Hubble constant. Consult the paper [5] !
CMBR Spectrum: Calculate the photon number density n
CMBR
contained in the Planck
spectrum and compare this number with the mean baryon number density n
B
in the present
Universe.
Observed galaxy redshift distributions: The overall luminosity function for galaxies can
be well approximated by means of a Schechter function (see e.g. Virgo cluster and other
nearby clusters)
(L) dL =

(L

/L)

exp(L/L

) d(L/L

) . (2.18)
L

is a characteristic luminosity,

the corresponding normalisation of the space density, and


1 the powerlaw index of the distribution. The number of galaxies observed in a redshift
bin dz with energy ux f in the range [f, f + df] and in the solid angle 4 is then given
by, for z 1,
d
2
N
dz df
= 4
_ _

0
(L/L

) d(L/L

) r
2
dr (r cz/H
0
) (L 4fr
2
) (2.19)
Show that the above integral is given by
d
2
N
dz df
=
4
L

_
c
H
0
_
5
z
4
(z
2
) , (2.20)
where = 4fc
2
/(H
2
0
L

) is a dimensionless constant. In a uxlimited survey, we observe


all galaxies with uxes above some minimum ux
dN
dz
=
_

f
min
d
2
N
dz df
df =
_
c
H
0
_
3
z
4
_

min
(z
2
) d, (2.21)
where
min
= 4f
min
c
2
/(H
2
0
L

).
Explain by simple arguments, why the observed distribution scales as z
4
, compare with
Fig. 2.19.
Plot the expected distribution dN/dz for the Schechter function and compare with observed
distributions for given ux limits.
Galaxy Correlation function: What is the meaning of the correlation function (r) for a
3D galaxy distribution ?
What is the correlation length r
0
? (see e.g. CfA or Las Campanas surveys)
How can one nd estimators for this correlation function ?
Spherical Harmonics: Look up the denition of spherical harmonics as used e.g. in quan-
tum mechanics. Explain the completeness of this functional basis.
Bibliography 41
Galaxy cluster gas: Explain the emission process behind the Xrays observed from galaxy
clusters. What processes heat the gas to Xray temperatures ?
Discuss the cooling times for this cluster gas (Virgo cluster e.g.) as a function of radius from
the cluster center.
Jeans equation: What is the Jeans equation for the dynamics in galaxies ?
How can one extract the mass distribution in a galaxy by using the Jeans equation ?
Bibliography
[1] Alpher, R.A., Herman, R.C.: 1948, Nature 162, 774-775
[2] P. Astier, J. Guy, N. Regnault et al.: 2005, The Supernova Legacy Survey: Mea-
surement of
M
,

and w from the First Year Data Set, astroph/0510447


[3] Bennett, C.L. et al.: 2003, First Year Wilkinson Microwave Anisotropy Probe
(WMAP) Observations: Preliminary Maps and Basic Results, ApJS 148, 1.
[4] Freedman, W.L. et al.: 1994, Nature 371, 757
[5] Freedman, W.L. et al.: 2001, Final Results from the Hubble Space Telescope Key
Project to Measure the Hubble Constant, ApJ 553, 47.
[6] Gamow, G.: 1948, Nature 162, 680-682
[7] Healpix Homepage: http//www.eso.org/healpix
[8] Hinshaw, G. et al.: 2003, First Year Wilkinson Microwave Anisotropy Probe
(WMAP) Observations: The Angular Power Spectrum, ApJS 148, 135
[9] Hu, W., Dodelson, s.: 2002, Cosmic Microwave Background Anisotropies, Ann.
Rev. Astron. Astrophys. 40, 171.
[10] Hubble, E.: 1929, A relation between distance and radial velocity among extra-
galactic nebulae, Proc. Nat. Acad. Sci. (USA) 15, 168
[11] Penzias, A.A., Wilson, R.W.: 1965, Astrophysical Journal 142, 419-421 (July)
[12] Tonry, J.L. et al.: 2003, Cosmological Results from High-z Supernovae, ApJ 594, 1
[13] Tegmark, M.: 1996, ApJ 470, 81
[14] Vogeley, M.S., Park, C., Geller, M.J., Huchra, J.P.: 1992, Largescale clustering of
galaxies in the CfA Redshift Survey, Astrophsy. J. Letters 391, L5
Part II
Relativistic World Models
3 The Relativistic Cosmos
3.1 Relativity and Cosmology
General relativity is the currently accepted theory of gravitation having been introduced by
Einstein in 1916, superceding the newtonian theory. It plays a major role in astrophysics in
situations involving strong gravitational elds, for example the study of neutron stars, black
holes, and gravitational lensing. The theory also predicts the existence of gravitational radi-
ation, which manifests itself by the transfer of energy due to a changing gravitational eld,
for example that of a binary pulsar. General relativity therefore also provides the theoretical
foundation for the subject of cosmology, in which one studies the structure and evolution of
the Universe on the largest possible scales.
The nal steps to the theory of general relativity were taken by Einstein and Hilbert at
almost the same time. Both had recognised aws in Einsteins October 1914 work and a
correspondence between the two men took place in November 1915. How much they learnt
from each other is hard to measure but the fact that they both discovered the same nal form
of the gravitational eld equations within days of each other must indicate that their exchange
of ideas was helpful.
On the 18th November Einstein made a discovery about which he wrote For a few days I
was beside myself with joyous excitement. The problem involved the advance of the perihe-
lion of the planet Mercury. Le Verrier, in 1859, had noted that the perihelion (the point where
the planet is closest to the sun) advanced by 38 per century more than could be accounted
for from other causes. Many possible solutions were proposed, Venus was 10% heavier than
was thought, there was another planet inside Mercurys orbit, the sun was more oblate than
observed, Mercury had a moon and, really the only one not ruled out by experiment, that New-
tons inverse square law was incorrect. This last possibility would replace the 1/d
2
by 1/d
p
,
where p = 2+ for some very small number. By 1882 the advance was more accurately known,
43 per century. From 1911 Einstein had realised the importance of astronomical observations
to his theories and he had worked with Freundlich to make measurements of Mercurys orbit
required to conrm the general theory of relativity. Freundlich conrmed 43 per century in
a paper of 1913. Einstein applied his theory of gravitation and discovered that the advance
of 43 per century was exactly accounted for without any need to postulate invisible moons
or any other special hypothesis. Of course Einsteins 18 November paper still does not have
the correct eld equations but this did not affect the particular calculation regarding Mercury.
Freundlich attempted other tests of general relativity based on gravitational redshift, but they
were inconclusive.
Also in the 18 November paper Einstein discovered that the bending of light was out by
a factor of 2 in his 1911 work, giving 1.74. In fact after many failed attempts (due to cloud,
war, incompetence etc.) to measure the deection, two British expeditions in 1919 were to
46 3 The Relativistic Cosmos
conrm Einsteins prediction by obtaining 1.980.30 and 1.610.30.
On 25 November Einstein submitted his paper The eld equations of gravitation which
give the correct eld equations for general relativity. The calculation of bending of light and
the advance of Mercurys perihelion remained as he had calculated it one week earlier.
Five days before Einstein submitted his 25 November paper Hilbert had submitted a paper
The foundations of physics which also contained the correct eld equations for gravitation.
Hilberts paper contains some important contributions to relativity not found in Einsteins
work. Hilbert applied the variational principle to gravitation and attributed one of the main
theorems concerning identities that arise to Emmy Noether who was in G ottingen in 1915.
No proof of the theorem is given. Hilberts paper contains the hope that his work will lead to
the unication of gravitation and electromagnetism.
Immediately after Einsteins 1915 paper giving the correct eld equations, Karl Schwarzschild
found in 1916 a mathematical solution to the equations which corresponds to the gravitational
eld of a massive compact object. At the time this was purely theoretical work but, of course,
work on neutron stars, pulsars and black holes relied entirely on Schwarzschilds solutions
and has made this part of the most important work going on in astronomy today.
The starting point for the application of Einsteins theory to cosmology is what is termed
cosmological principle (sometimes also called the Copernican principle):
Viewed on sufciently large distance scales, there are no preferred directions or pre-
ferred places in the Universe.
Stated simply, this principle means that averaged over large enough distances, one part of the
Universe looks approximately like any other part. In this sense, the Earth is not a preferred
location in the Universe the physical laws tested in our labs should apply to all positions in
the Universe.
In this Section we shortly describe the essential elements of Einsteins theory of gravity
and derive the most general form of isotropic world models. This leads to spaces of constant
curvature. We then discuss observational aspects for such world models: the origin of redshift,
distance measurements and number counting.
Natural Scaling: From Hubbles constant H
0
we derive a characteristic scale, often called
the present Hubble radius of the Universe
R
H

c
H
0
= 4286 Mpc , H
0
= 70
km
s Mpc
. (3.1)
In comparison to the extensions of superclusters with dimensions of about 50 Mpc, this is
about a hundred times larger. Only gravitational forces can act over such large distances. The
Hubble constant also means a fundamental timescale
t
H
=
1
H
0
= 13.97 10
9
h
1
0.7
yr . (3.2)
This is a measure for the age of the Universe. All distance scales are given in terms of R
H
and all timescales in terms of t
H
.
It is therefore clear that only a relativistic theory of gravity can deal with such huge scales
and times. Newtonian gravity with its instantaneous action for forces is not compatible with
the basic ideas of Special Relativity, where the speed of light is the characteristic propagation
time for all interactions between matter. The General Theory of Relativity brings together
Special Relativity and gravitational forces. This theory is a geometric description of gravity.
3.2 Einsteins Vision of Gravity 47
A central idea, the equivalence principle, suggests that motion attributed to a gravitational
force is to interpreted as freefall motion in a curved spacetime. Since the motion is due to
curvature, all objects are affected equivalently. This curvature may be thought as geometrys
deviation from at Minkowski spacetime. In the language of differential geometry, spacetime
can be described as a 4dimensional real differentiable manifold M on which a metric g
ab
with Lorentzian signature is dened. This metric allows one to meassure spacetime separa-
tions in a coordinateindependent manner. This metric tensor eld describes gravity since all
measurable properties can be derived from it. In the at space limit, all equations reduce to
those of Special Relativity. Since this theory is well tested in the Solar system and even in
a few compact binary systems, we have good reasons to believe that this theory also models
gravity in the cosmos correctly.
3.2 Einsteins Vision of Gravity
Modern cosmology begins with Hubbles observation that the universe of galaxies is expand-
ing. A theoretical basis for this observation has been given by Einsteins theory of gravity,
more than 10 years earlier. In modern terms, Einsteins theory of gravity is a gauge theory
with the Lorentz group as the gauge group which is operating in the tangent space. Gravity is
therefore modeled by means of an afne connection of a Lorentzian manifold.
1
In this Sec-
tion, we summarize all the elements necessary to understand the geometry of the Friedmann
Universe and of its generalisations, such as the perturbed Friedmann Universe or Brane Cos-
mology. However, it is not the purpose of this Section to introduce all concepts in sufcient
depth, for this attend a lecture on General Relativity.
3.2.1 The Concept of SpaceTime
Special relativity showed that the absolute space and time of Newtonian physics could be only
an approximation to their true nature. However, the special theory of relativity is incapable
of explaining gravity because SR assumes the existence of inertial frames; it does not explain
how inertial frames are to be determined. Machs principle, which states that the distribu-
tion of matter determines space and time, suggests that matter is related to the denition of
inertial frames, but Mach never elucidated any means by which this might happen. General
relativity attacks this problem and in so doing, discovers that gravity is related to geometry.
The equivalence principle is the fundamental basis for the general theory of relativity. The
strict equivalence between gravity and inertial acceleration means that freefalling frames are
completely equivalent to inertial frames. In general relativity, (GR) it is spacetime geometry
that determines freefalling (inertial, geodesic) worldlines, telling matter how to move. Mat-
ter, in turn, tells spacetime how to curve. Geometry is related to matter and energy through
Einsteins equation. The metric equation provides a general formalism for the spacetime in-
terval in general geometries, not just the Minkowski (at) spacetime of special relativity (SR).
Matter and energy determine inertial frames, but within an inertial frame there is no inuence
1
A modern introduction into General Relativity can be found in the textbooks by Carroll [2] and Hobson et
al. [3]. The latter one does include the basic concepts for Cosmology (Friedmann models and Ination). A more
mathematically oriented treatment is given by Straumann [7]. This textbook also includes a complete overview for
modern differential geometry of Riemannian manifolds (theory of tensor elds, afne connections, curvature and
pforms).
48 3 The Relativistic Cosmos
Figure 3.1: In 1916 Albert Einstein published the fundamental paper uber die relativistische
Theorie der Gravitation.
by any outside matter. Thus Machs principle is present more in spirit than in actuality in the
general theory of relativity.
The Concept of a Metric
To introduce the concept of a metric, let us consider Euclidean 2dimensional space with
Cartesian coordinates x, y. A parametrized cureve (x(t), y(t)) begins at t
1
and ends at t
2
.
The length of the curve is given by
s =
_
ds =
_
_
dx
2
+dy
2
=
_
t
2
t
1
_
x

2
+y

2
dt . (3.3)
Here, ds =
_
dx
2
+dy
2
is the line element. The square of the line element, also called the
metric, is then given as
ds
2
= dx
2
+dy
2
. (3.4)
For this representation, we also can use polar coordinates (r, ) with the expression for the
metric
ds
2
= dr
2
+r
2
d
2
. (3.5)
3.2 Einsteins Vision of Gravity 49
In a similar manner, in 3dimensional Euclidean space, the metric is given by
ds
2
= dx
2
+dy
2
+dz
2
, (3.6)
in Cartesian coordinates, and
ds
2
= dr
2
+r
2
(d
2
+ sin
2
d
2
) (3.7)
in spherical coordinates.
Einstein I: Minkowski space M
4
has to be generalized to a general curved
pseudoRiemannian manifold (M, g) with metric tensor eld g such that SpaceTime is
locally still Minkowskian, i.e. the tangent space T
p
M = M
4
.
The notion of an event is fundamental in relativity. An event is characterized by its position
x and its time t. An event p is given by the 4 coordinates (t, x, y, z) in the 4dimensional
SpaceTime. Already in Special Relativity, time and space appear as an entity. Two neighbor-
ing events (t, x, y, z) and (t + dt, x + dx, y + dy, z + dz) have then a distance ds which is
determined by the metric of Minkowski space, x
0
= ct,
ds
2
= c
2
dt
2
dx
2
dy
2
dz
2
=

dx

dx

. (3.8)
This distance is invariant against Lorentz transformations. We often say the spacetime of SR
is at, since the resulting curvature vanishes.
A suitable tool to picturize a spacetime is to use spacetime diagrams (Fig. 3.2). The
time-axis is running vertically, and space is running horizontally. In Minkowski space, the
lightcones have a constant openening angle of 90 degrees. In a curved spacetime this may
change.
Gravity cannot be included into Special Relativity despite many desperate attempts to do
this 100 years ago. Einstein postulated therefore that Minkowski space is only realized locally
in a 4DpseudoRiemannian manifold (M, g). This means strictly speaking, Minkowski space
M
4
is the tangent space T
p
M= M
4
at each event p of the spacetime M. Einstein had the idea
that the effects of gravity are expressed in terms of a generalized Minkowski metric element
of the form
ds
2
=
3

,=0
g

(x) dx

dx

, (3.9)
which gives the distance between neighboring events in M. The ensemble of all events
parametrized by local coordinates {t, x
i
} is called the SpaceTime. The metric tensor g is
a secondrank symmetric tensor, which in general depends on the events. ds now measures
the proper time of timelike curves x

()
= s/c =
_
_
g

dx

d
dx

d
d. (3.10)
Examples of Simple SpaceTimes
The Schwarzschild spacetime as the expression of the gravitational eld of nonrotating
stars
ds
2
= exp(2(r)) c
2
dt
2
exp(2(r)) dr
2
r
2
(d
2
+ sin
2
d
2
) . (3.11)
50 3 The Relativistic Cosmos
Figure 3.2: The light cones in Minkowski space are at. The timeaxis runs vertically, the
spatial axes horizontally. At each event we nd a forward and backward light cone. Photons (and
other massless particles) move along the light cone, while the trajectories of normal particles are
conned to the interior of the light cones. A detector can only measure photons which come in
from the backward light cone.
It has two independent functions, which only depend on the spherical radius r.
The gravitational eld of rapidly rotating stellar objects
ds
2
=
2
c
2
dt
2
R
2
(d dt)
2
exp(2
r
) dr
2
exp(2

) d
2
(3.12)
already has 5 independent functions depending now on the radius r and on , but not on
and t, = (r, ) etc. This line element contains an offdiagonal term g
0
, related to
the angular momentum of the star.
Cosmological spacetimes
ds
2
= c
2
dt
2
R
2
(t) d
2
(3)
, (3.13)
where d
(3)
is the metric of a 3space of constant curvature. The essential degree of
freedom is the expansion factor R(t) which scales all spatial lengths (see later on). The
simplest example is the stretching of Minkowski space (called a at Universe)
ds
2
= c
2
dt
2
R
2
(t)
_
dx
2
+dy
2
+dz
2
_
, (3.14)
often written in spherical coordinates (t, r, , ) as
ds
2
= c
2
dt
2
R
2
(t)
_
dr
2
+r
2
(d
2
+ sin
2
d
2
)
_
. (3.15)
All of the above spacetimes have some high degree of symmetries.
3.2 Einsteins Vision of Gravity 51
3.2.2 Gravity is an Afne Connection on SpaceTime
In order to compare tangent spaces at neighboring events, one needs a connection on the
manifold M. As in Riemannian geometry, this connection is required to be metric, so that the
corresponding Christoffel symbols are uniquely given by derivatives of the metric elements.
This is a basic postulate of Einsteins theory of gravity one could construct more general
theories of gravity which include torsion.
Physically speaking, we associate observers e
a
, a = 0, 1, 2, 3, i.e. an orthonormal tetrad
(or Vierbein eld), satisfying
2
g(e
a
, e
b
) =
ab
, (3.16)
where is the at Minkowskian metric with signature (+), or (+++). An observer
is a global orthonormal basis eld in the tangent space of each event p, where e
0
is timelike
and e
i
(i = 1, 2, 3) are spacelike. One could also construct null tetrads in order to dene
the geometry of the spacetime. The dual elements of e
a
is a basis of the cotangent space
T

p
, denoted by
a
, satisfying
a
(e
b
) =
a
b
. They dene the metric g =
ab

a

b
. The
denition of these observer elds is not unique, since any observer derived by means of a local
Lorentz transformation is also an observer
e
a
|
x
=
b
a
(x) e
b
|
x
, (x)
T
(x) = . (3.17)
These are Lorentz transformations operating in the tangent space of each event.
As an example we consider static observers in the Schwarzschild metric (3.11). Such an
observer is given by the following tetrad
e
0
= exp()
t
, e
r
= exp()
r
, e

=
1
r

, e

=
1
r sin

. (3.18)
It is then clear that they satisfy g(e
a
, e
b
) =
ab
, where
ab
is the Minkowski metric. The dual
basis is a basis of oneforms
a
with
a
(e
b
) =
a
b

0
= exp() dt ,
r
= exp() dr ,

= r d ,

= r sin d. (3.19)
We now consider a satellite which is orbiting the central star in the equatorial plane of the
Schwarzschild spacetime with 4velocity u

given by
u = u
t
(
t
+

) , g(u, u) = 1 . (3.20)
= u

/u
t
= d/dt is the angular velocity of the satellite (Keplerian e.g.) as measured
by xed stars. The Lorentz transformation between the static observer e
a
and the satellite
observer e
a
is then given by a boost transformation with 3velocity V = r sin / and
Lorentz factor
S
= 1/

1 V
2
, where
2
= 1 2GM/r (c = 1). is the redshift factor
between a static observer at radius r and innity. This provides us the Lorentz transformation
e
0
=
S
(e
0
+V e

) (3.21)
e
r
= e
r
(3.22)
e

= e

(3.23)
e

=
S
(V e
0
+e

) . (3.24)
The trajectory of this observer, with tangent vector e
0
is now a helical path in spacetime.
2
In the following, the convention for indices is as follows: greek indices are related to local coordinate systems,
latin indices a, b, c, ... mark observer elds, latin indices i, k, l, ... specify spatial components.
52 3 The Relativistic Cosmos
The Concept of a Connection
A connection is now dened as a linear mapping between the tangent space at the event x and
the tangent space at a neigboring event displaced by dx. It is sufcient to dene this mapping
for an arbitrary basis e
a
of the tangent space

e
a
e
b
=
c
ab
e
c
=
c
b
(e
a
) e
c
, (3.25)
with the additional properties for any function f and any vector eld X

fX
e
b
= f
X
e
b
(3.26)

X
(fe
b
) = f
X
e
b
+ (X.f)e
b
. (3.27)
Thus, the oneforms dened as

b
a
=
b
ca

c
(3.28)
are called connection oneforms. They are identical with the Christoffel symbols, if the basis
in the tangent space is the natural basis implied by the coordinate system

. (3.29)
Remember that the rst index in the Christoffel symbols is a oneform index, the second
is a matrix index.
From the duality between tangent and cotangent space we nd then

a
=
a
b
(X)
b
(3.30)
for any vector eld X. From this denition we nd the covariant derivative for any vector
eld X = X
a
e
a
X = e
a
(dX
a
+
a
b
X
b
) , (3.31)
or in components with respect to an orthonormal basis

a
X
b
= e

a
X
b
,
+
b
c
(e
a
)X
c
. (3.32)
Similarly, for a oneform =
a

a
we have
=
b
(d
b

a
b
) , (3.33)
or in components

b
= e

b,

c
b
(e
a
)
c
. (3.34)
When we use the coordinate basis of the chosen chart, the covariant derivatives of vector elds
and oneforms are given in the wellknown form

= X

,
+

(3.35)

=
,

. (3.36)
3.2 Einsteins Vision of Gravity 53
The covariant derivative for vector elds and oneforms can now be extended to arbitrary
tensor elds, in general, by requiring that the operation of satises the Leibniz rule when
acting on tensor products
(S T) = S T+S T. (3.37)
In this sense, the covariant derivative of a 2tensor eld T

is given as follows

= T

,
+

, (3.38)
and similarly for a 2tensor eld A

by means of

= A
,

. (3.39)
Parallel Transport and Geodesics
A connection on a vector bundle (here the tangent space) species then the notion of parallel
transport along curves in the manifold. Let be a curve on the manifold, and X a vector
eld dened on M. A vector eld is called autoparallel along , if


X = 0 . (3.40)
In coordinates, we have
X = X

, =
dx

ds

, (3.41)
and therefore


=
_
dX

ds
+

dx

ds
X

. (3.42)
The vector eld is autoparallel if
dX

ds
+

dx

ds
X

= 0 . (3.43)
For any curve (s) and X
0
in the tangent space T
(0)
Mwe nd a unique vector eld X(s)
given along (s) with the initial condition X(0) = X
0
. This operation is called the parallel
displacement of a vector eld along a curve (s).
A curve is called a geodesic, if the tangent eld is autoparallel along (s). According
to the above analysis, this means
d
2
x

ds
2
+

dx

ds
dx

ds
= 0 . (3.44)
Geodesics are the trajectories of freely falling bodies in the gravitational eld given by the
afne connection. A satellite e.g. will move on geodetic curves in the gravitational eld of the
Earth, planets move on the geodetic curves in the solar gravitational eld.
54 3 The Relativistic Cosmos
Gravity is a Metric Connection
So far, the concept of a metric and the concept of the connection are independent of each
other. Each Riemannian manifold, however, carries a particular connection which is uniquely
associated with the metric. For this, we say:
An afne connection is said to be a metric connection if the parallel transport along any
smooth curve in the manifold preserves the inner product.
One can then prove that this statement is equivalent to g = 0, which is equivalent to the
Ricci identity
X.g(Y, Z) = g(
X
Y, Z) +g(Y,
X
Z) . (3.45)
With the condition that torsion vanishes,

X
Y
Y
X [X, Y ] = 0 , (3.46)
we can write the above equation as
X.g(Y, Z) = g(
Y
X, Z) +g([X, Y ], Z) +g(Y,
X
Z) . (3.47)
[X, Y ] denotes the Lie bracket for vector elds (see next Section). With cyclic permutation of
the vector elds we obtain
Y.g(Z, X) = g(
Z
Y, X) +g([Y, Z], X) +g(Y,
X
Z) (3.48)
Z.g(X, Y ) = g(
X
Z, Y ) +g([Z, X], Y ) +g(X,
Z
Y ) . (3.49)
Now we add the rst and third equation and subtract the second one to get
2g(
Z
Y, X) = X.g(Y, Z) +Y.g(Z, X) +Z.g(X, Y )
g([Z, X], Y ) g([Y, Z], X) +g([X, Y ], Z) . (3.50)
For the fundamental vector elds X =
k
, Y =
j
and Z =
i
the Lie bracket vanishes,
[
i
,
j
] = 0 and g(
i
,
j
) = g
ij
, which means that
2g(

j
,
k
) = 2
m
ij
g
mk
=
k
g
ji
+
j
g
ik
+
i
g
kj
(3.51)
or
g
mk

m
ij
=
1
2
(g
jk,i
+g
ik,j
g
ij,k
) . (3.52)
With the inverse metric g
ij
, we now get the famous expression for the LeviCivita connection

m
ij
=
1
2
g
mk
(g
ki,j
+g
kj,i
g
ij,k
) . (3.53)
This afne connection is therefore uniquely determined by derivatives of the metric ten-
sor and is therefore called metric connection, or LeviCivita connection.
For any pseudoRiemannian manifold there is then a unique afne connection such that it
is (i) torsionfree and (ii) metric. This particular connection is usually called the LeviCivita
connection, or pseudoRiemannian connection.
Einstein II: It is now one of the fundamental postulates of Einsteins theory of gravity
that gravity is related to the LeviCivita connection of the Lorentzian manifold. This
means in particular that there is no torsion associated with gravity.
3.2 Einsteins Vision of Gravity 55
Strong Principle of Equivalence
Since the Lorentz connection transforms inhomogeneously as
= (x)
1
(x) (d)
1
(x) (3.54)
for any Lorentz transformation between local observers,

a
=
a
b
(x)
b
, we always can nd
locally an observer system such that
p
= 0, i.e. the connection can be transformed away just
locally, but not globally. This is not the case for the curvature !
The weak principle of equivalence states that effects of gravitation can be transformed
away locally by suitably accelerated frames of reference (by going to local Minkowskian co-
ordinates). We can formulate, however, a much stronger requirement, the socalled
Einstein III: strong principle of equivalence, which states that any physical interaction
(other than gravitation) behaves in a local inertial frame as if gravitation were absent.
E.g. Maxwells equations will have their familiar forms as in SR.
The strong principle of equivalence allows us to extend any physical law that is expressed
in a covariant way to curved SpaceTime. Ordinary derivatives are just replaced by covariant
ones.
3.2.3 Differential Forms on SpaceTime
Differential forms are extremely helpful concepts in direct calculation. A zeroform is a scalar
function. The oneforms
a
dened above are the basis elements of the cotangent space, its
components are the components of covariant vectors. A general oneform A can always
be written as A = A

dx

= A
a

a
. The vector potential of classical electrodynamics is the
standard example. A new operation introduced when one works with forms is called the wedge
product. If x and y are coordinates, then dx and dy are oneforms, and dx dy = dy dx
is called a twoform. An example of a pform is
A =
1
p!
A
...
dx

dx

... dx

, (3.55)
where A
...
is a completely antisymmetric tensor with p indices. In fact, the set of pforms
in a ndimensional manifold is a vector space
p
of dimension n!/p!(n p)! (see Table 3.1).
In 4 dimensions we have one zeroform, 4 oneforms (basis in the cotangent space), 6 2forms
(the Farady tensor e.g.), 4 3forms (currents) and only one 4form (volumeform). Formally,
the wedge product of a pform with a qform is given by the alternating operator
= A( ) . (3.56)
The exterior derivative d takes a pform into a (p + 1)form, e.g. a oneform
d = d(

dx

) =
,
dx

dx

=
1
2
(
,

,
) dx

dx

. (3.57)
In general, for a pform A given by
A = A
...
dx

dx

... dx

(3.58)
56 3 The Relativistic Cosmos
Forms 0 1 2 3 4
dim = 3 1 3 3 1
dim = 4 1 4 6 4 1
Table 3.1: Number of linearly independent pforms for D = 3 and D = 4.
the exterior derivative is given by its local expression
dA = dA
...
dx

dx

... dx

=
A
...
x

dx

dx

dx

... dx

. (3.59)
With this explicit denition, one can show
d(A B) = dA B + (1)
p
A dB (3.60)
d(dA) = 0 (3.61)
for pform A and a qform B.
One can also dene an antiderivation i which makes a (p 1)form out of a pform
dened as
(i
V
)(V
1
, ..., V
p1
) = (V, V
1
, ..., V
p1
) , (3.62)
i.e. just by contraction with the rst index. With the Farady tensor F we can e.g. build the
oneform E = i
V
F, in components E

= V

. This operation is called the inner product


of V with . Applying both operations, the inner product and the exterior derivative, leaves
the degree of a pform invariant
L
X
= d i
X
+i
X
d (3.63)
is equivalent to the Lie derivative on pforms.
The Lie derivative L is given by its action on functions
L
X
f = X.f = df(X) , (3.64)
its action on vector elds
L
X
Y = [X, Y ] = (X

,
Y

,
) e

, (3.65)
and the Leibniz rule for the compatibility with higher rank tensors
L
X
(S T) = L
X
S T +S L
X
T . (3.66)
From the last property, we can derive for a oneform and a vector eld Y
L
X
( Y ) = L
X
((Y )) = (L
X
) Y + (L
X
Y ) . (3.67)
Writing out in components, we have
X

)
,
= (L
X
)

(L
X
Y )

, (3.68)
or making use of the Liebracket
(L
X
)

= X

(
,
Y

,
)

(X

,
Y

,
)
= (X

,
+

,
) Y

. (3.69)
Since this last equation is valid for any vector eld Y , we conclude
(L
X
)

= X

,
+

,
. (3.70)
3.2 Einsteins Vision of Gravity 57
3.2.4 Curvature of SpaceTime
A physical theory of gravity also requires some dynamical evolution for the connection. This
is usually formulated in terms of the curvature associated with the connection. The calculation
of the Riemann tensor is therefore one of the major tasks when dealing with specic space-
times. For this purpose we denote by X(M) the space of all (smooth) vector elds on the
manifold M.
Conventionally, the torsion elds T are dened as bilinear mappings T : X(M)
X(M) X(M) on the set of all vector elds on the manifold
T(X, Y )
X
Y
Y
X [X, Y ] . (3.71)
Curvature is dened as a trilinear mapping R : X(M) X(M) X(M) X(M)
R(X, Y )Z
X
(
Y
Z)
Y
(
X
Z)
[X,Y ]
Z . (3.72)
The components R
a
bcd
of this vector eld denes the Riemann tensor which has four indices.
These quantities obviously satisfy the antisymmetry conditions
T(X, Y ) = T(Y, X) , R(X, Y ) = R(Y, X) , (3.73)
as well as
T(fX, gY ) = fg T(X, Y ) (3.74)
R(fX, gY )hZ = fghR(X, Y )Z (3.75)
for any functions f, g and h.
Since torsion T(X, Y ) and curvature R(X, Y ) are antisymmetric tensors, they naturally
dene corresponding twoforms
T(X, Y ) = T
a
(X, Y ) e
a
(3.76)
R(X, Y )e
b
=
a
b
(X, Y ) e
a
. (3.77)
The exterior derivatives of the basic oneforms
a
and of the connection forms satisfy
Cartans structure equations
T
a
= d
a
+
a
b

b
(3.78)

a
b
= d
a
b
+
a
d

d
b
. (3.79)
The wedge operator denotes the exterior products for pforms. The 2form is the curvature
2form which gives, when expressed locally,

a
b
=
1
2
R
a
bcd

c

d
(3.80)
the components of the Riemann tensor R
a
bcd
in orthonormal coordinates. Similarly, we have
4 torsion twoforms
T
a
=
1
2
T
a
bc

b

c
. (3.81)
58 3 The Relativistic Cosmos
For the proof of Cartans structure equations, we use the above denition of torsion. Writ-
ten as oneforms, this means
T
a
(X, Y ) e
a
=
X

Y

Y

X
[X, Y ]
=
X
(
b
(Y )e
b
)
Y
(
b
(X)e
b
)
a
([X, Y ])e
a
=
_
X.
a
(Y ) Y.
a
(X)
a
([X, Y ])
_
e
a
+
_

a
(Y )
b
a
(X)
a
(X)
b
a
(Y )
_
e
a
= d
a
(X, Y )e
a
+ (
a
b

b
)(X, Y )e
a
. (3.82)
The proof of the second structure equation is similar. Written as a twoform, this means

a
c
(X, Y ) e
a
=
X

Y
e
c

X
e
c

b
c
([X, Y ])e
b
=
X
(
b
c
(Y )e
b
)
Y
(
b
c
(X)e
b
)
a
c
([X, Y ])e
a
=
_
X.
a
c
(Y ) Y.
a
c
(X)
a
c
([X, Y ])
_
e
a
+
_

b
c
(Y )
a
b
(X)
b
c
(X)
a
b
(Y )
_
e
a
= d
a
c
(X, Y )e
a
+ (
a
b

b
c
)(X, Y )e
a
. (3.83)
Local expressions: In local coordinates, a metric connection is expressed in terms of the
Christoffel symbols

=
1
2
g

_
g
,
+g
,
g
,
_
(3.84)
such that the connection form is given in a local coordinate basis as

dx

, (3.85)
and therefore
d

,
dx

dx

=
1
2
(

,
) dx

dx

. (3.86)
Also,

dx

dx

=
1
2
(

) dx

dx

(3.87)
Accordingly, Cartans second structure equation is equivalent to the conventional denition of
the Riemann tensor in local coordinates
R

,
+

. (3.88)
3.2.5 Curvature and Einsteins Equations
Another consequence of the afne connection is an additional symmetry of the Riemann tensor
g(R(X, Y )Z, U) = g(R(X, Y )U, Z) (3.89)
g(R(X, Y )Z, U) = g(R(Z, U)X, Y ) . (3.90)
3.2 Einsteins Vision of Gravity 59
The Riemann tensor is the fundamental entity for the construction of the eld dynamics. It
satises the following essential symmetries which are important for the concrete calculation
R
a
bcd
= R
a
bdc
(3.91)
R
abcd
= R
bacd
(3.92)
R
abcd
= R
cdab
. (3.93)
The rst property results from the fact that curvature is a twoform, the second one that cur-
vature is an element of the Lie algebra of the Lorentz group, and the third one gives a funda-
mental relation between spacetime indices and intrinsic indices (metric condition). This last
property follows from the cyclic identity for a torsionfree connection
R(X, Y )Z +R(Z, X)Y +R(Y, Z)X = 0 , (3.94)
or in components
R
abcd
+R
adbc
+R
acdb
= 0 . (3.95)
Making use of the antisymmetry in the rst and second pair of indices, we nd
R
abcd
= R
adbc
R
acdb
= R
dabc
+R
cadb
= (R
dcab
+R
dbca
) (R
cbad
+R
cdba
)
= 2R
cdab
+ (R
bdca
+R
bcad
)
= 2R
cdab
R
badc
= 2R
cdab
R
abcd
. (3.96)
Hence
R
abcd
= R
cdab
. (3.97)
In total, the Riemann tensor has 36 components, while the last symmetry reduces it to 20
independent components. The Riemann tensor of apcetimes has 20 independent compo-
nents. Astrophysical spacetimes have usually many symmetries such that the total number of
independent components is drastically reduced. In comparison, the metric tensor has only 10
independent components, i.e. only half of the components of the Riemann tensor are due to
the metric, or the Ricci tensor R
ab
, while the other 10 components are hidden in the Weyl
tensor.
The Riemann tensor is now used to construct the Ricci tensor, R
bd
= R
a
bad
. For the
Schwarzschild spacetime (3.11) e.g. we get the following expressions for the Ricci tensor
R
00
= R
r
0r0
+R
2
020
+R
3
030
(3.98)
R
rr
= R
0
r0r
+R
2
r2r
+R
3
r3r
(3.99)
R
22
= R
0
202
+R
r
2r2
+R
3
232
(3.100)
R
33
= R
0
303
+R
r
3r3
+R
2
323
(3.101)
with all other components vanishing. With the Ricci scalar R = R
a
a
as the trace of the Ricci
tensor we now can construct the Einstein tensor
G
ab
R
ab

1
2
Rg
ab
. (3.102)
60 3 The Relativistic Cosmos
The Einstein tensor is symmetric, G
ab
= G
ba
, and divergencefree, G
a
b;a
= 0 (due to the
Bianchi identity).
Einstein IV: Einstein postulated that the tensor G
ab
couples to the matter content of
the Universe
G
ab
=
8G
c
4
T
ab
, (3.103)
where T
ab
is the symmetric energymomentum tensor of all matter in the Universe (par-
ticles, baryons, galaxies, photons, neutrinos and quantum elds). As a consequence of the
above properties, the divergence of the energymomentum tensor vanishes identically
T
ab
;b
= 0 . (3.104)
Einsteins equations can be derived from the action
A =
1
16G
_
R

g d
4
+
_
L
matter
(, )

g d
4
x (3.105)
where L
matter
is the Lagrangian density for matter depending on some variables denoted
collectively as , c = 1, since for any domain D of spacetime [7]

_
D
R

g d
4
x =
_
D
G

g d
4
x. (3.106)
The variation of this action with respect to will lead to the equation of motion for matter,
L
matter
/ = 0, while the variation of the action with respect to the metric tensor g leads to
Einsteins equations
3
R


1
2
Rg

= 16G
L
matter
g

8GT

. (3.107)
Here, T

= 2L
matter
/g

is the energymomentum tensor of matter elds.


On the Cosmological Constant
Let us now consider a new matter action L

matter
= L
matter
/(8G), where is a real
constant. The equation of motion for the matter does not change under this transformation,
since is constant. But the action now picks up an extra term proportional to , which can
be written in two different ways,
A =
1
16G
_
R

g d
4
x +
_
_
L
matter
(, )

8G
_

g d
4
x
=
1
16G
_
(R 2)

g d
4
x +
_
L
matter
(, )

g d
4
x (3.108)
and Einsteins equations get modied. This simple manipulation has many backdrops in the-
oretical Physics. It can be interpreted in different manners:
3
see any textbook on General Relativity
3.2 Einsteins Vision of Gravity 61
The rst interpretation is based on the rst line of the above equations, it treats as a
shift in the matter Lagrangian, which in turn will lead to a shift in the matter Hamiltonian.
This could be thought of as a shift in the zero point energy of the matter system. Such
a constant shift in energy does not affect the dynamics of matter, while gravity picks up
an extra contribution in the form of a new term Q

in the energymomentum tensor


R


1
2
R

= 8G(T

+Q

) , Q

=

8G

. (3.109)
The second line in Eq (3.108) can be interpreted as a gravitational eld, described by
the Lagrangian of the form L
grav
(1/G)(R 2), interacting with matter. In this
interpretation, gravity is described by two constants, the Newtons constant G and the
cosmological constant . It is then natural to modify the left hand side of Einsteins
equations in the form of
R


1
2
R

= 8GT

. (3.110)
In this interpretation, the spacetime is curved even in the absence of matter, T

= 0,
since the left hand side does not admit at spacetimes as solutions.
It is even possible to consider a situation where both effects can occur. If gravitational
theories are in fact described by the Lagrangian of the form (R 2), then there is
an intrinsic cosmological constant in nature, just as there is a Newtonian constant G
in nature. If the matter Lagrangian contains energy densities which change due to the
dynamics, then L
matter
can pick up constant shifts during dynamical evolution. For this
we consider a scalar eld with the Lagrangian
L

= (1/2)

V () , (3.111)
which has the energymomentum tensor
T

_
1
2

V ()
_
. (3.112)
For eld congurations which are constant (e.g. at the minimum of the potential V ), this
contributes an energymomentum tensor T

= V (
min
)

, which has exactly the same


form as a cosmological constant. It is then the combination of these two effects of very
different nature which is relevant and the source will be
T
e

= [V (
min
) + /(8G)] g

. (3.113)

min
can change during the dynamical evolution, leading to a timedependent cosmo-
logical constant.
The term Q

in Einsteins equations behaves very pecularly compared to the energy


momentum tensor of normal matter. Q

is in the form of an energymomentum


tensor of an ideal uid with energy density

and pressure P

. Obviously, either the


pressure or the energy density of this uid must be negative.
Such an equation of state, P = , also has another important implication in GR. The
relative acceleration between two geodesics, g, satises in GR the following equation
g = 4G( + 3P) . (3.114)
62 3 The Relativistic Cosmos
The source of this relative acceleration between geodesics is + 3P and not alone. This
shows, as long as +3P > 0, gravity remains attractive, while +3P < 0 leads to repulsive
forces. A positive cosmological constant therefore leads to repulsive gravity.
Remark: Gravity as a Gauge Theory
We now have the means to compare the formalism of connections and curvature in Rieman-
nian geometry to that of gauge theories in particle physics. In both situations, the elds of
interest live in vector spaces which are assigned to each point in spacetime. In Riemannian
geometry the vector spaces include the tangent space, the cotangent space, and the higher
tensor spaces constructed from these. In gauge theories, on the other hand, we are concerned
with internal vector spaces. The distinction is that the tangent space and its relatives are inti-
mately associated with the manifold itself, and were naturally dened once the manifold was
set up; an internal vector space can be of any dimension we like, and has to be dened as an
independent addition to the manifold. In math lingo, the union of the base manifold with the
internal vector spaces (dened at each point) is a ber bundle, and each copy of the vector
space is called the ber (in perfect accord with our denition of the tangent bundle).
Nongravitational interactions are described nowadays in terms of gauge theories. In this
sense, Maxwells theory is a U(1) gauge theory resulting from local phase transformations on
quantum elds, (x) exp(i(x)) (x). The vector potentials A

(x) are the connection


coefcients, and the Faraday tensor F = (1/2)F

dx

dx

is the corresponding curvature.


Strong interaction is a SU(3) gauge theory, where the internal space is dened by the color
space each fermion can carry a specic color. The gauge elds A
b
a
are then the local
expressions of a connection oneform = A

dx

with values in the Liealgebra of SU(3).


In this case, we have 8 connection elds A

(x), = 1, ..., 8, corresponding to the 8 gluon


elds of strong interaction.
In this sense, the gauge transformations for gravity are the local Lorentz transformations
operating between different observers at the same events in spacetime. We have 6 connection
elds A

(x), = 1, ..., 6, i.e. in total 24 connection coefcients. Note, however, that the
dynamics proposed by Einstein is different from the YangMills dynamics of nonAbelian
gauge theories.
3.3 General Relativity is the Correct Theory of Gravity
Most of the tests for Einsteins theory of gravity have been done for stellar objects, such as the
Sun or neutron stars. In good aprroximation, stars are spherical objects and the gravitational
eld is given in terms of spherically symmetric metric elements.
4
Das einfachste metrische Feld wird von einem kugelsymmetrischen Stern erzeugt. In
diesem Falle reduziert die hohe Symmetrie (Kugelsymmetrie) die m oglichen metrischen Ko-
efzienten auf zwei wesentliche Funktionen g
00
(r) und g
rr
(r)
ds
2
= exp(2(r)) dt
2
exp(2(r)) dr
2
r
2
(d
2
+ sin
2
d
2
) . (3.115)
Die Einsteinschen Feldgleichungen, welche die Kr ummung der RaumZeit mit dem Ma-
terieinhalt verkn upfen, bestimmen uber Differentialgleichungen diese beiden Funktionen ein-
4
A detailed analysis of all these tests is not the topic of the present lecture, see e.g. any lecture on GR or on
Relativistic Astrophysics.
3.3 General Relativity is the Correct Theory of Gravity 63
deutig, r > R

,
exp(2(r)) = exp(2(r)) = 1
2GM
c
2
r
, g

= r
2
, g

= r
2
sin
2
. (3.116)
Diese Metrik ist als SchwarzschildMetrik eines Sterns bekannt. Sie enth alt nur die Gesamt-
masse M des Sterns als freien Parameter. Die Gr oe
R
S
=
2GM
c
2
= 3 km
M
M

(3.117)
ist der der Masse M. Dies war die erste L osung der von Einstein 1915 postulierten Feldgle-
ichungen. Das Gravitationsfeld der Sonne mu durch ein solches metrisches Feld beschrieben
werden, obschon die Abweichungen vom achen Raum an der Ober ache der Sonne nur
R
S
/R

10
6
betragen. Bei Neutronensternen sind diese Abweichungen jedoch schon
betr achtlich, R
S
/R

1/3.
Figure 3.3: Die Lichtkegel werden in der Nahe eines Schwarzen Lochs verzogen: sie werden
auf das Zentrum zu gerichtet. Der Zylinder wird durch die Zeitentwicklung des Horizontes
erzeugt (1 Dimension unterdruckt).
Gravitational Redshift
Betrachten wir z.B. einen Beobachter mit r = const, = const und = const und stellen
uns die Frage, wie die Zeit seiner Uhr sich zur Koordinatenzeit t verh alt (die Zeit im Un-
64 3 The Relativistic Cosmos
endlichen). d = ds/c mit die Eigenzeit des Beobachters in einem lokalen Inertialsystem.
Deshalb gilt
d =
_
1
2GM
c
2
r
dt . (3.118)
Wie die Metrik zeigt, ist t die im Unendlichen gemessene Zeit. Gegen uber Unendlich scheint
deshalb eine Uhr im Gravitationsfeld langsamer zu gehen. Dies k onnen wir auch dadurch
ausdr ucken, dass Photonen im Gravitationsfeld eines Sterns rotverschoben werden

A
=
d
A
d
B
=

g
00
(A)
g
00
(B)
(3.119)
ergibt das Verh altnis der Frequenzen eines Photons an verschiedenen Stellen A und B im
Gravitationsfeld. Beobachten wir etwa die Emission einer Linie mit Wellenl ange

auf der
Ober ache eines kompakten Sterns, so erhalten wir

B
=
1
_
1
2GM

c
2
R

. (3.120)
Die Spektrallinien eines kompakten Sterns werden deshalb rotverschoben
z =

B

=
1
_
1
2GM

c
2
R

1
GM
c
2
R

. (3.121)
Dies ist die sog. gravitative Rotverschiebung. F ur Weie Zwerge hat man Rotverschiebun-
gen z 10
4
10
5
gemessen (Sirius B), bei Neutronensternen w urden sich Rotverschiebun-
gen z 0.2 ergeben und bei Schwarzen L ochern z = . Das Schwarze Loch hat gerade
einen Radius R

= R
S
(Abb. 3.3).
PostKeplerian Effects
Apart from gravitational redshift, three other general relativistic effects are observable in the
solar system and are nowadays of principal importance for the calculation of ephemerids of
planets:
The perihelion precession for the Mercury orbit by 43 arcsec per century (Fig. 3.4);
Light deection on the solar surface by

=
4GM

c
2
R

= 1.75 (3.122)
The Shapiro timedelay for signals propagating in the solar system. This effect is due to
longer propagation of signals in the space curved by the Sun compared to a propagation
far away from the solar surface.
General Relativity predicts the bending of light by gravity, gravitational time dilation and
length contraction, gravitational redshifts and blueshifts, the precession of Mercurys orbit,
and the existence of gravitational radiation. All these effects have been measured, although
3.4 Isotropic SpaceTimes 65
Figure 3.4: Relativistische Periheldrehung im Zweikorperproblem. Fur die Merkurbahn
betragt diese postNewtonsche Periheldrehung nur 43 Bogensekunden pro Jahrhundert (hier
stark ubertrieben).
gravitational radiation has been observed only indirectly via the decay of the orbits of binary
pulsars. The LIGO project is an attempt to build a giant Michelson-Morley type of interfer-
ometer to detect gravitational radiation directly. Two interferometers have been built, each
one with perpendicular light-carrying vacuum pipes 5 kilometers long.
The relativistic periastron shift and Shapiro timedelay are essential effects used in astron-
omy to determine the exact pulse arrival times for radio pulses emitted by pulsars in compact
binary systems (see Camenzind 2006).
3.4 Isotropic SpaceTimes
The observed high degree of isotropy of the CMBR implies strong constraints on possible
world models. In a rst part we have to regain the 3+1 aspects of a spacetime.
3.4.1 Slicing of SpaceTime
Any spacetime (M, g) can be decomposed by means of the diffeomorphism
: M I (3.123)
where is considered as a spatial part (3D space) and I as the timeow. The entire spacetime
consists then of a set of slices given by some intrinsic metric. This decomposition is called
66 3 The Relativistic Cosmos
Figure 3.5: In a gravitational lense, the gravitational eld of a galaxy e.g. deects the photon
paths so that multiple images can occur.
Figure 3.6: Photon trajectories are strongly affected by the gravitational eld of a rotating Black
Hole. The Black Hole is bombarded by laser photons along the equatorial plane. Photons with
low impact parameters are captured by the horizon.
the 3 + 1split of spacetime.
The tangent vector eld
t
can now be decomposed into a part normal to the slice and one
along the slice

t
= n +

. (3.124)
3.4 Isotropic SpaceTimes 67
Figure 3.7: Slicing of a spacetime into a sequence of 3spaces with metric . The slicing
denes the lapse function and the shift vector

.
n is a normal vector to the slice
t
, is called lapse function, and

which is parallel to
t
is called shift vector. We dene now a particular observer given by the vector eld n and 3
orthonormal vector elds e
(3)
i
in the slice
e
0
= n =
1

_
, e
i
= e
(3)
i
. (3.125)
Similarly, we can introduce a basis of oneforms adapted to the slice

0
= dt ,
i
=
i
+
i
dt . (3.126)
From a direct calculation, one can verify

a
(e
b
) =
a
b
, (3.127)
and the metric is expressed as
g =
ab

a

b
. (3.128)
With this we can calculate
g(
t
,
t
) =
0
(
t
)
0
(
t
) +
ik

i
(
t
)
k
(
t
)
=
2
+
i

k
(3.129)
g(
t
, e
i
) =
0
(
t
)
0
(e
i
) +
ik

i
(
t
)
k
(e
i
) =
i
. (3.130)
68 3 The Relativistic Cosmos
Spacetimes around stellar objects are asymptotically at, i.e. Minkowskian far from the
surface of the star. Cosmological spaces are not asymptotically at, but they should have
timeslices which are required to be homogeneous and isotropic. It is then a well known fact
in differential geometry that a Riemannian manifold of dimension n > 2 which is required to
be isotropic about each point is necessarily of constant curvature.
In the cosmos we require the existence of a unique timeow which is associated with the
clusters of galaxies. In terms of this time t, spacetime can be sliced (i, k = 1, 2, 3)
ds
2
=
2
dt
2

ik
(t, x)(dx
i
+
i
dt)(dx
k
+
k
dt) . (3.131)
The shift vectors
i
have to vanish according to our time choice. The 3tensor
ik
is then the
metric of the slices
t
for const time t. These slices are 3D Riemannian manifolds which are
now required to be isotropic around each point.
In a cosmological spacetime (M, g), there will be a family of preferred world lines
representing the average motion of matter at each point (these represent the histories of
clusters of galaxies, with associated fundamental observers). Their 4velocity is
u

=
dx

d
(3.132)
where is the proper time measured along the world lines. At recent times this velocity is
taken to be the 4velocity dened by the vanishing of the CMBR dipole. There is precisely
one such 4velocity which will set this dipole to zero. It is usually assumed that this is the
same as the average 4velocity of matter in a suitably sized volume.
3.4.2 Isotropic Riemannian Spaces
Observations tell us that there is a special ow vector eld u for which the Universe always
looks isotropic. This vector eld is associated with the mass centers of galaxy clusters (Virgo,
Coma etc). The ow vector eld of individual galaxies always includes some particular mo-
tions due to the connement in a cluster.
Isotropy of SpaceTimes
Isotropy applies at some specic point in the space, and states that the space looks the same no
matter what direction you look in. More formally, a manifold M is isotropic around a point p
if, for any two vectors V and W in T
p
M, there is an isometry of M such that the pushforward
of W under the isometry is parallel with V (not pushed forward).
Homogeneity is the statement that the metric is the same throughout the space. In other
words, given any two points p and q in M, there is an isometry which takes p into q. Note that
there is no necessary relationship between homogeneity and isotropy; a manifold can be ho-
mogeneous, but nowhere isotropic, or it can be isotropic around a point without being homo-
geneous (such as a cone, which is isotropic around its vertex but certainly not homogeneous).
On the other hand, if a space is isotropic everywhere then it is homogeneous. (Likewise if it is
isotropic around one point and also homogeneous, it will be isotropic around every point.)
5
The existence of this global symmetry has then essential consequences:
5
This can be formulated more mathematically for the case of a spacetime: A spacetime is said to be isotropic in
each of its events p with respect to a given vector eld u with g(u, u) = 1, if the set ISO
p
of all isometries which
leave invariant the point p is a subset of the rotational group SO(3)(V
p
). All these transformations operate in the
space orthogonal to u.
3.4 Isotropic SpaceTimes 69
The vector eld u has to be orthogonal to the slices
t
, i.e. the corresponding shift vector
elds

have to vanish,

= 0 for all events.
The integral curves of the vector eld u are geodesics,
u
u = 0. This has the conse-
quence that the distance between two slices
t
1
and
t
2
is independent of time. This
leads to the introduction of a preferred cosmic time t such that
ds
2
= dt
2
R
2
(t)
ik
dx
i
dx
k
. (3.133)
3.4.3 Spaces of Constant Curvature
If a manifold Mis isotropic, then its sectional curvature K
p
(E) only depends on p and not on
the plane E in T
p
M, and, by Schurs Lemma it has constant sectional curvature. This means
that the hypersurfaces
t
are spaces of constant curvature.
As we have seen, for a given connection on M curvature is given by a tensor of rank
(1,3), i.e. by a linear mapping of T
p
M T
p
M T
p
M T
p
M dened by the relation (see
Sect. 3.2.4)
R(u, v)w =
u

v
w
v

u
w
[u,v]
w. (3.134)
u, v and w are vector elds on M, [u, v] = L
u
v is the Liebracket of two vector elds. This
can also be written in terms of a fourth vector eld z as
R(u, v, w, z) g(R(u, v)w, z) = R

. (3.135)
This tensor is antisymmetric in both pairs of indices and satises the Bianchi identity.
The sectional curvature with respect to a plane E in T
p
M is given by
K
p
(E) = R(e
1
, e
2
, e
1
, e
2
) , (3.136)
when (e
1
, e
2
) is a basis of E. If the sectional curvature is constant for all planes E in T
p
M
and for all events p, this denes a space of constant curvature.
Schurs Lemma: Isotropy around each point leads to a homogeneous space. Or ex-
actly: If the sectional curvature K
p
(E) in a Riemannian space of dimension n 3 is
independent of E for each point p, then K
p
(E) is also independent on p, i.e. M is a space
of constant curvature.
Mathematically, a Riemannian space that has constant curvature must have a Riemann curva-
ture tensor of the following form
R
abcd
= K(g
ac
g
bd
g
ad
g
bc
) (3.137)
In this equation K is a constant sometimes simply called the curvature of the space. In spaces
of constant curvature, the spaces are qualitatively different, depending on K > 0, K = 0 and
K < 0. The Ricci tensor has then the form
R
bd
= g
ac
R
abcd
= K(n 1)g
bd
, (3.138)
and the Ricci scalar
R = R
a
a
= K n(n 1) . (3.139)
70 3 The Relativistic Cosmos
The curvature 2form assumes the simple form

ab
= K
a

b
. (3.140)
We now specialize to 3space and contract this to get the Riccitensor
R
bd
= g
ac
R
abcd
= 2Kg
bd
. (3.141)
Hence, in the 3Dsubspace, the Ricci tensor is proportional to the metric tensor.
3.4.4 FriedmannRobertsonWalker (FRW) SpaceTimes
Conformal Coordinates: Riemann spaces of constant curvature are conformally at,
i.e. there is a coordinate system such that
6
=
1

i
(dx
i
)
2
. (3.146)
In a next step we prove that in a space of constant curvature one can always nd local
coordinates x
i
such that the metric is expressible as
7
=

i=1,n
(dx
i
)
2
(1 +K
2
/4)
2
,
2
=

i
(x
i
)
2
. (3.157)
6
In order to prove this we use the fact that a manifold with dimension n 3 is at if and only if the Weyl tensor
vanishes. The Riemann tensor can be decomposed into a part given by the Ricci tensor and a second part given by
the Weyl tensor C
abcd
R
abcd
= C
abcd

1
n 2
_
g
ab
R
cd
+ g
bd
R
ac
g
bc
R
ad
g
ad
R
bc
_
+
1
(n 1)(n 2)
(g
ab
g
cd
g
ad
g
bc
)R. (3.142)
This tensor has the same symmetries as the Riemann tensor, but in addition its Riccipart vanishes
g
bd
C
abcd
= 0 . (3.143)
The deeper differential geometric meaning of the Weyl tensor is the conformal invariance, i.e. its invariance against
transformations of the form g exp((x)) g. Therefore, a metric is called conformally at if there is a transfor-
mation of the form
g
ab
exp()
ab
. (3.144)
This means that the Weyl tensor vanishes and that the forms
C
ab
=
1
2
C
abcd

c

d
(3.145)
identically vanish in a conformally at space. One can aslo show the opposite: if the Weyl tensor of a space vanishes,
then this space is conformally at.
7
For this we consider the metric of the form of Eq (3.146) and dene an orthonormal frame
i
= dx
i
/ with
=

i

i

i
. For the exterior derivative we obtain
d
i
=

,j

2
dx
i
dx
j
=
,j

i

j
=
i
k

k
. (3.147)
The last equality follows from the rst Cartan equation. Since
dg
ij
=
ij
+
ji
= 0 , (3.148)

ij
are antisymmetric. For this reason, we have the solution

ij
=
,j

i
+
,i

j
(3.149)
3.5 The FRW World Models 71
For a 3space we may express this line element in the form of
d
2
(3)
=
1
(1 +K
2
/4)
2

i=1,3
(dx
i
)
2
. (3.158)
We can use the 3D forms of the line element that we have constructed to form the full 4D
cosmological line element. In 4D we have
ds
2
= c
2
dt
2
R
2
(t) d
2
(3)
. (3.159)
R(t) is a general scaling function of time t. The set of all these spaces denes the Friedmann
RobertsonWalker (FRW) spacetimes. One can then always normalize the constant k =
1, 0.
3.5 The FRW World Models
In standard textbooks, the FRW model is given in a somewhat different coordinate system
ds
2
= c
2
dt
2
R
2
(t)
_
dr
2
1 kr
2
+r
2
(d
2
+ sin
2
d
2
)
_
. (3.160)
From the second Cartan equation we derive the curvature

ij
= d
ij
+
m
i

m
j
= (
,jm

m

i
+
,im

m

j
)
,i

,m

m

j
+ (
,m

i
+
,i

m
) (
,j

m
+
,m

j
) . (3.150)
By combining different terms we obtain

ij
=
,jm

m

i
+
,im

m

j

,m

,m

i

j
= K
i

j
. (3.151)
This is only possible if
,im
= 0 for i = m. is therefore a linear combination of functions f(x
i
), =

i
f(x
i
).
By inserting this into the curvature we get

ij
=
_
(f

i
+ f

j
)

m
(f
2
)
_

i

j
(3.152)
From this we nd
f

i
+ f

j
=
_

m
(f
2
) + K
_
/ (3.153)
Since the rhs is independent of i and j, we must have f

i
= f

j
= const, i.e. these are quadratic functions
f
i
(x
i
) = ax
2
i
+ b
i
x
i
+ c
i
. Therefore =
2
/4 + 1. By inserting into the curvature, we nd = k. Therefore,
we have found the solution
= 1 +
K
4

2
(3.154)

i
=
dx
i
1 + K
2
/4
(3.155)

ij
=
K
4
(x
i

j
x
j

i
) . (3.156)
The global classication of Riemannian manifolds with constant curvature is a delicate subject.
72 3 The Relativistic Cosmos
For this we make the substitution
r =

1 +K
2
/4
, dr =
r

d
K
2
r
2
d (3.161)
With this we can verify that
dr
2
1 kr
2
=
d
2
(1 +K
2
/4)
2
. (3.162)
Therefore the above metric can be expressed in the form
Figure 3.8: 3spaces of constant curvature.
ds
2
= dt
2

R
2
(t)
(1 +k
2
/4)
2
_
d
2
+
2
(d
2
+ sin
2
d
2
_
. (3.163)
For k = 1, we can show that the r = 1 singularity is just a coordinate singularity by
changing to the specially suited coordinates involving the new coordinate dened in terms
of r as
r = sin (3.164)
The singularity is therefore eliminated since
dr = cos d =
_
1 r
2
d (3.165)
The timeslice is then given by
d
2
(3)
= R
2
[d
2
+ sin
2
(d
2
+ sin
2
d
2
)] . (3.166)
Simalarly, the open case, k = 1, is obtained by r = sinh such that
dr = cosh d =
_
1 +r
2
d. (3.167)
The corresponding timeslice has the form
d
2
(3)
= R
2
[d
2
+ sinh
2
(d
2
+ sin
2
d
2
)] , (3.168)
3.6 The Origin of the Cosmological Redshift 73
with the following ranges
0 < , 0 , 0 2 . (3.169)
Flat space, k = 0, can trivially be expressed in terms of these coordinates
d
2
(3)
= R
2
(t)
_
d
2
+
2
(d
2
+ sin
2
d
2
)
_
. (3.170)
Summary
We have found three equivalent formulations (i.e. coordinate systems) for a FriedmanRobertson
Walker Universe
in quasispherical coordinates (t, r, , )
ds
2
= c
2
dt
2
R
2
(t)
_
dr
2
1 kr
2
+r
2
(d
2
+ sin
2
d
2
)
_
. (3.171)
in conformal coordinates (t, , , )
ds
2
= c
2
dt
2

R
2
(t)
(1 +k
2
/4)
2
_
d
2
+
2
(d
2
+ sin
2
d
2
)
_
. (3.172)
in hyperspherical coordinates (t, , , )
ds
2
= c
2
dt
2
R
2
_
d
2
+S
2
()(d
2
+ sin
2
d
2
)
_
, (3.173)
where S() = {sin , , sinh()} depending on the curvature.
The only degree of freedom is the expansion factor R(t), which has to be determined by
Einsteins equations.
3.6 The Origin of the Cosmological Redshift
The FRW model explains in a natrual way the cosmological redshift. For this we may assume
that each galaxy is given by its comoving coordinates (r, , ). Let us consider a galaxy with
coordinates (r
1
, , ) emitting photons at time t
1
. Photons follow null geodesics, ds
2
= 0
(Fig. 3.9). Thus
c dt =
Rdr

1 kr
2
. (3.174)
Since r decreases for light propagating towards us, the minus sign is appropriate. For photons
detected at time t
0
we then have
_
t
0
t
1
c dt
R(t)
=
_
r
1
0
dr

1 kr
2
. (3.175)
74 3 The Relativistic Cosmos
Figure 3.9: Worl lines in an expanding universe: the middle line represents the oberver, neigh-
boring world lines are associated with neighboring galaxies. The backward light cone is locally
similar to the at space limit, is however strongly curved in the early universe. Photons detected
by observers move along the backward light cone and are emitted at the intersections between
the light cone and the trajectory of the galaxy.
Let us consider wave crests, emitted at t
1
and t
1
+ t
1
_
t
0
+t
0
t
1
+t
1
c dt
R(t)
=
_
r
1
0
dr

1 kr
2
. (3.176)
For R(t) roughly constant over t
0
, we obtain by subtracting both equations
c t
0
R(t
0
)

c t
1
R(t
1
)
= 0 . (3.177)
Therefore,
c t
0
c t
1
=

1

0
=

0

1
=
R(t
0
)
R(t
1
)
= 1 +z , (3.178)
with z as the redshift. The wavelength of the photons get stretched by propagating through
the universe, R(t
0
) > R(t
1
). This is the natural explanation of the redshift detected by Hub-
ble in 1929.
3.7 The Luminosity Distance and the HubbleLaw
Photons represent collisionless particles and must be treated in terms of phasespace distri-
bution functions f(x; p) with 3momentum p. We consider a stream of particles propagating
3.7 The Luminosity Distance and the HubbleLaw 75
r
e
a(t
o
)
r

=

r
e
t
o
r

=

r
e
dt
e D
r

=

0
Figure 3.10: Observations in an expanding Universe. The observers world line is at r = 0, the
emitters world line at r = r
e
which is stretched to a(t
0
)r
e
at present time. The emitter object
has a physical extension D.
freely in a FRWspacetime. At some time t, a comoving observer nds dN particles in a
proper volume dV , all having momenta in the range [ p, p + d
3
p]. The phasespace distri-
bution function f for the particles is now dened by the relation dN = f dV d
3
p. At a
later instant t + t, the proper volume of the particles would have increased by a factor
[R(t + t)/R(t)]
3
, whereas the volume in the momentum space would be redshifted by
the same factor [R(t)/R(t + t)]
3
. The total phase volume occupied by the particles does
therefore not change during free propagation. Because the number of particles dN is also
conserved without interactions, the phasespace distribution function f is conserved along
the streamline.
Photon momenta are characterized by the energy E = h and a direction n on the unit
sphere, p = (h/c)n. The momentum space volume is therefore d
3
p = p
2
dp d
2
d d,
so that the photon number is given by
f

=
dN

d
3
xd
3
p
=
dN

[cdt
e
dA
e
][
2
e
d
e
d
e
]
=
dN

[cdt
0
dA
0
][
2
0
d
0
d
0
]
. (3.179)
d is the solid angle around around the direction of propagation n, and d
3
x cdt dA, where
dA is the area normal to the direction of propagation. The subscript e represents the process
of emission and 0 the process of detection. This shows that
f

=
dN

dt dA
2
d d
(3.180)
is invariant along the streamline. In Astronomy, instead of f

the radiation eld is character-


76 3 The Relativistic Cosmos
ized in terms of the specic intensity I

= I

(x; n) dened as
h dN

= I

dt dAd d. (3.181)
This shows that I

/
3
is invariant along streamlines.
The quantity I

will have units erg s


1
cm
2
Hz
1
Steradian
1
. The specic energy
density u

= (4/c)I

will have the dimensions erg cm


3
Hz
1
. The energy ux in the
radiation scales therefore as
F
(1+z)
(z) = F

(0) (1 +z)
3
. (3.182)
The total ux of radiation is obtained by integration over the full spectrum and varies therefore
as
F =
_

0
F

d (1 +z)
4
. (3.183)
The energy density in the CMBRscales therefore as u (1+z)
4
, which is compatible with the
StephanBoltzmann law u = aT
4
. Radiation elds of the form I

=
3
G(/T) will retain
the spectral shape under the expansion of the Universe, since the temperature is redshifted as
T R
1
. As a consequence, a Planckian spectrum remains a Planckian spectrum under the
expansion (see COBE results for the CMBR).
The Luminosity Distance
Cosmological observations are based on electromagnetic radiation of sources (galaxies, quasars,
GRBs) at high redshift. Let an observer located at r = 0 receive at time t
0
radiation of a source
located at r = r
1
, emitted at earlier time t
1
(Fig. 3.9). Let Lt
1
the total energy emitted
by a galaxy per unit of time, at time t
1
< t
0
. This energy will be detected by an observer in
the timeinterval t
0
= t
1
(R(t
0
)/R(t
1
)), and the energy will be redshifted by the amount
R(t
1
)/R(t
0
). For isotropic emission, we can calculate the radiation ux of a 2sphere with
radius r
1
. For t = const and r = const the lineelement is given by
ds
2
(2)
= r
2
R
2
(d
2
+ sin
2
d
2
) . (3.184)
This is the line element of an Euclidean 2sphere with radius rR. At time t
0
, the light emitted
by a galaxy G is distributed on a 2sphere with surface A = 4r
2
1
R
2
(t
0
). The ux observed
in a detector corresponds to the emitted energy by means of the relation
F t
0
=
Lt
1
4r
2
1
R
2
(t
0
)
R(t
1
)
R(t
0
)
, (3.185)
i.e. the observed ux is given by
F =
t
1
t
0
R(t
1
)
R(t
0
)
L
4R
2
(t
0
)r
2
1
=
L
4R
2
0
r
2
1
_
R(t
1
)
R(t
0
)
_
2
. (3.186)
Here, we used relation (3.178) between the time interval in the observed system and the time
interval of the emitter.
8
This determines the expression for the bolometric luminosity
F
bol
=
L
4r
2
1
R
2
(t
0
) (1 +z)
2
=
L
4 d
2
L
(3.187)
8
The stretching of timescales has been seen in the light curves of Supernovae: the observed decay time of the
light curve is longer than the intrinsic decay time, t
0
= (1 + z)
SN
.
3.7 The Luminosity Distance and the HubbleLaw 77
For this reason, the quantity
d
L
(z) r
1
R(t
0
)(1 +z) (3.188)
is called the luminosity distance of an object with redshift z. This distance is used in the
classical distance modulus
m
bol
M
bol
= 5 log[d
L
/pc] 5 . (3.189)
The HubbleLaw
The Hubblelaw is a general consequence of the FRW model for the Universe and does not
depend on the particular expansion law. For this we make a Taylor expansion around the
present state of the Universe at time t
0
R(t) = R(t
0
[t
0
t])
= R(t
0
)

R(t
0
t) +
1
2

R(t
0
t)
2
+O(t
3
)
= R
0
_
1 +H
0
(t t
0
)
1
2
q
0
H
2
0
(t t
0
)
2
+O(t
3
)
_
. (3.190)
Here we introduced the Hubble parameter H and the deceleration parameter q
0
H
0
= (

R/R)
t
0
(3.191)
q
0
= [

R/(RH
2
)]
t
0
. (3.192)
A dot always corresponds to differentiation with respect to cosmic time t. These notions are
valid at any cosmic time. We will show in the following that H
0
is indeed equivalent to the
Hubble constant. Since the expansion factor is equivalent to redshift, 1 + z = R
0
/R(t), the
above formula can also be written as
1 +z =
_
1 +H
0
(t t
0
)
1
2
q
0
H
2
0
(t t
0
)
2
+O(t
3
)
_
1
1 +H
0
(t
0
t) +
1
2
q
0
H
2
0
(t t
0
)
2
+H
2
0
(t t
0
)
2
+O(t
3
) . (3.193)
This means for the redshift
z = H
0
(t
0
t) +
_
1 +
1
2
q
0
_
H
2
0
(t t
0
)
2
+O(t
3
) . (3.194)
For not too early times, i.e. H
0
|t t
0
| 1, or z 1, this can be inverted to give
t
0
t =
1
H
0
_
z (1 +q
0
/2) z
2
_
+O(z
3
) . (3.195)
t
0
t is the time since the emission of a photon at time t (socalled lookback time). It is
important to note that these relations only depend on the presentday values H
0
and q
0
of
the Hubble and decelration parameters.
78 3 The Relativistic Cosmos
Since photons propagate along null geodesics, d = 0,
ds
2
= dt
2
R
2
(t)
dr
2
1 kr
2
= 0 , (3.196)
we nd for the relation between the propagation time and distance
_
t
0
t
cdt
R(t)
=
_
r
0
dr

1 kr
2
r . (3.197)
This integral gives
rR
0
=
_
t
0
t
R
0
R(t)
dt =
1
R
0
_
t
0
t
(1 +z) dt
=
1
R
0
_
t
0
t
_
1 +H
0
+
_
1 +
1
2
q
0
_
H
2
0

2
_
d
= c(t
0
t)
_
1 +
H
0
2
(t
0
t) +O([H
0
(t t
0
)]
2
)
_
. (3.198)
Using the above relation between time and redshift, equation (3.195), this yields
rR
0
=
c
H
0
_
z (1 +q
0
/2)z
2
+
1
2
z
2
+O(z
3
)
_
=
c
H
0
_
z
1
2
(1 +q
0
)z
2
+O(z
3
)
_
. (3.199)
This shows that the metric distance rR
0
is given by the Hubble radius c/H
0
and that curvature
effects enter in quadratic order of redshift.
By using the general expression for the luminosity distance, d
L
= r
1
R
0
(1+z), we recover
Hubbles law upto second order in the redshift, which is typically valid for redshifts z < 0.1,
i.e. upto distances of about 500 Mpc (see Fig. 2.4),
d
L
= d
L
(z) =
c
H
0
_
z +
1
2
(1 q
0
)z
2
+O(z
3
)
_
. (3.200)
Hubbles law would be correct even in second order of redshift in a decelerating Universe with
q
0
= 1.
The luminosity distance for redshifts z > 0.1 depends on the particular FRW model,
i.e. on the expansion law R(t). For a de Sitter model driven by vacuum energy, R(t) =
R
i
exp(H(t t
i
)), one nds (q
0
= 1)
d
dS
L
(z) =
c
H
z(1 +z) . (3.201)
Luminosity distances at highredshift behave in general quadratic with the redshift, and not
linear.
As a fundamental fact we have found that the Hubble constant H
0
is just the present
expansion velocity of the Universe
H
0
=
_

R
R
_
0
. (3.202)
3.8 The Hubble Constant 79
In Astronomy, the radiation ux f is usually expressed in terms of magnitudes
m = 2.5 log f +const , (3.203)
which can be handled by the absolute magnitude
M
bol
= 2.5 log
L
L

4.75 . (3.204)
L

= 3.90 10
26
Watt is the solar luminosity. This gives rise to the distance modulus
DM = mM = 5.0 log[d
L
/10 pc] . (3.205)
In suitable units, this can be written then as, [H
0
] = km/s/Mpc, [cz] = km/s,
DM = 48.35 5 log h + 5 log(z) + 1.086(1 q
0
)z . (3.206)
3.8 The Hubble Constant
As we have discussed in Sect. 2.2, the determination of the Hubble constant is one of the key
tasks in observational cosmology. The various methods which lead to still somewhat different
values for H
0
are summarized in Fig. 3.11 which is due to Tammann [8]. A key element in
this procedure is the calibration of the distance modulus for the Virgo cluster.
Method (mM)
Virgo
Galaxy type
Cepheids 31.52 0.20 S
Novae 31.46 0.40 E
Globular Clusters 31.67 0.15 E
Tully-Fisher 31.58 0.24 S
D
n
31.85 0.19 S0, E
Average 31.66 0.09 21.5 0.9 Mpc
Table 3.2: Distance modulus to the Virgo cluster according to Tammann [8].
3.9 The Apparent Angular Width of Galaxies
Another important measurable quantity is the apparent angular diameter of a galaxy . What
is the angular extension of a galaxy at redshift one, or what is the angular width of a jet with
length of 1 Mpc at redshift 5 ? For this consideration we may assume that the corners are
given by (
1
,
1
) and (
1
+d,
1
). From this we obtain the true width of the source
ds
2
= r
2
1
R
2
(t
1
) (
1
)
2
= D
2
(3.207)
80 3 The Relativistic Cosmos
Figure 3.11: Distance ladder for the measurement of the Hubble constant according to Tam-
mann. These calibrators do not include SN Ia.
with the physical diameter D of a galaxy. Therefore, the angular width is given by

0
=
D
r
1
R(t
1
)
=
D(1 +z)
r
1
R(t
0
)
. (3.208)
This can be wrtitten in terms of the luminosity distance

0
=
D
d
A
=
D(1 +z)
2
d
L
, d
A
= d
L
(z)/(1 +z)
2
. (3.209)
This demonstrates that the classical result D/d
L
is wrong for higher redshifts. Since d
L

c/H
0
, the overall angular width of cosmological objects is always of the order of D/R
H
.
3.10 Number Counts in the Expanding Universe 81
For a galaxy cluster e.g. this means
cl
10
3
. In a de Sitter model, the apparent angular
width of an object will be constant at high redshifts

dS
0
=
D
d
A
(z)
=
D
R
H
(1 +z)
2
z(1 +z)

D
R
H
. (3.210)
This is in fact the minimal angular width, since in a realistic model the angular width will
slightly increase with redshift beyond redshift one. A typical disk galaxy with extension of 30
kpc will subtend an angle of 710
6
rad, which amounts to 1.4 arcseconds. A galaxy cluster
with an extension of 3 Mpc will subtend a minimal angle of about 2 arcminutes.
3.10 Number Counts in the Expanding Universe
The number of cosmological sources in a volume section at a point P down the null cone is
dN = d
2
A
d
0
[n(k
a
u
a
)]
P
dy . (3.211)
n is the number density of radiating sources (galaxies, quasars, GRBs) per unit proper volume
in a section of bundle of light rays converging towards the observer and subtending a solid
angle d at the observers position, d
A
is the area distance of this section from the observers
point of view (also called angular distance), u
a
is the observers 4velocity, k
a
the tangent
vector of the null rays and y the afne parameter distance down the light rays (Fig. 3.12). n
is measured in the rest frame of the counted sources. This expression is metric independent
and therefore valid for all cosmological models (rst derived by Ellis 1973).
The above form is written in unobservable variables. Since the redshift is given by
1 +z =
(u
a
k
a
)
P
(u
a
k
a
)
Observer
(3.212)
the number density can be expressed in terms of redshift
dN = d
2
A
d
0
[n]
P
(1 +z)(k
a
u
a
)]
O
dy . (3.213)
The minus sign is chosen because we are interested in incoming light rays. k is a null vector
satisfying
k
a
k
a
= 0 , k
b
k
a;b
= 0 . (3.214)
Therefore k
a
u
a
= 1 and
dN = d
2
A
d
0
(1 +z) n(y) dy . (3.215)
n and d
A
are functions of z, as well as y, and therefore we refer to the quantity dN/dz instead
of dN/dy.
FRW Model: The distance along the line of sight corresponding to a redshift interval dz is
given by
dl = c dt = c
dt
dR
dR
dz
dz =
dz
1 +z
cR

R
=
dz
1 +z
d
H
(z) , (3.216)
82 3 The Relativistic Cosmos
Figure 3.12: Number counts of objects detected in an expanding universe.
where d
H
(z) = c/H(z) is the Hubbleradius at redshift z. In a given cosmological model
we can determine dl/dz = d
H
/(1 + z). From this we obtain the proper volume element at
redshift z
dV =
d
H
(z) dz
1 +z
d
2
A
d =
R
2
0
r
2
em
(z)d
H
(z)
(1 +z)
3
dz d. (3.217)
Thus the number count of galaxies per unit solid angle and per redshift interval should be
given by
dN
dz d
= n(z)
dV
dz d
= n(z)
R
2
0
r
2
em
(z)d
H
(z)
(1 +z)
3
= n(z) d
H
(z)
d
2
L
(z)
(1 +z)
5
. (3.218)
To compute this quantity explicitly, we need the luminosity distance d
L
(z) as well as the
Hubble radius as functions of redshift, which will be determined by means of the Friedmann
equation.
3.11 Slightly Inhomogeneous World Models 83
3.11 Slightly Inhomogeneous World Models
The CMBR demonstrates that the present Universe is highly isotropic. In lowest order of
approximation, the CMBR temperature does not depend on directions on the Sky. The same
holds for the Hubble constant. One could, however, imagine of world models where the
expansion slightly depends on directions on the Sky, but the matter would still be uniformly
distributed. Such Universes are called homogeneous. These models played some role e.g. in
investigations of singularities.
The metric element can still be split in the 3+1 manner
ds
2
= dt
2

ik
dx
i
dx
k
. (3.219)
The manifolds M are then locally M = R G the product of the timeaxis with a group
space. Completely inhomogeneous world models are difcult to construct. They are treated
only in the limit of small perturbations (ripples of spacetime including gravitational waves).
An approach to this will be discussed in Chapter 8.
3.12 Summary
A consistent theory for the description of gravity in the Universe has to based on a curved
spacetime with a metric tensor g

which embodies the gravitational potentials.


Gravitational forces are related to the metric connection of this pseudoRiemannian
manifold. Torsion is absent in Einsteins theory of gravity.
Curvature of this manifold is generated via Einsteins equations by the matter distribution
in form of the energymomentum tensor for all kind of particles and elds.
The Copernican principle requires the cosmic spacetime to be homogeneous and isotropic.
The Friedman-Robertson-Walker Universe (FRW) is the most general model, its spatial
sections are either at, a 3sphere or a 3hyperboloid. The corresponding lineelement
is very simple
ds
2
= c
2
dt
2
R
2
(t)
_
dr
2
1 kr
2
+r
2
(d
2
+ sin
2
d
2
)
_
. (3.220)
(r, , ) are adapted to Sky coordinates and are called comoving coordinates, k = 0, 1.
In FRW, cosmological redshift of galaxies and quasars is a mere result of the stretching
of the 3space.
In FRW, the Hubble-law is a mere consequence of the expansion of the Universe, inde-
pendent of the particular form of the expansion.
The Hubble-law does not extend to redshifts beyond z 0.1. On scales with z >
0.1, luminosity distance, apparent angular widths and number counts will depend on the
specic world model.
The Hubble parameter is the relative expansion velocity and is therefore redshiftdependent.
84 3 The Relativistic Cosmos
3.13 Exercises
Distances in a de Sitter Model: Let us consider a particular FRWworld model given by
the expansion law R(t) = R
i
exp(H(t t
i
)), H = const, k = 0 (de Sitter model). Calculate
the deceleration parameter q;
the luminosity distance d
L
(z);
the angular width of an object with intrinsic diameter D as a function of redshift;
the observed number density dN/dz for a constant comoving density of objects.
Proper Volume: For the total cosmos, the number of galaxies in a spherical shell at radius
r and r +dr is then given by
dN =
4r
2
dr

1 kr
2
n(t) . (3.221)
The time t depends on r over the null cone condition
_
t
0
t
cdt

R(t

)
=
_
r
0
dr

1 kr
2
. (3.222)
The total number of galaxies upto the radius r
1
is then
N(r
1
) =
_
r
1
0
4r
2
n(t) dr

1 kr
2
. (3.223)
On the other hand, the volume V over a distance r follows for k = 1
V
+
(r) =
4
3
R
3
r
3
_
3
2
arcsinr
r
3

3
2

1 r
2
r
2
_
, (3.224)
and for k = 1 correspondingly
V

(r) =
4
3
R
3
r
3
_
3
2

1 +r
2
r
2

3
2
arsinhr
r
3
_
. (3.225)
For samll distances r 1 we get the Euclidean result with a small correction from curvature
V
k
(r) =
4
3
R
3
r
3
_
1 +
3
10
kr
2
+O(r
4
)
_
. (3.226)
Bibliography
[1] Camenzind, M.: 2007, Compact Objects in Astrophysics White Dwarfs, Neu-
tron Stars and Black Holes, SpringerVerlag (Heidelberg)
[2] Carroll, S.M.: 2004, Spacetime and Geometry, Addison Wesley (San Francisco)
(this is a modern introduction into the basics of General Relativity)
Bibliography 85
[3] Hobson, M.P., Efstathiou, G.P., Lasenby, A.N.: 2006, General Relativity An
Introduction for Physicists, Cambridge University Press
[4] Einstein, A.: 1917, Kosmologische Betrachtungen zur allgemeinen Rela-
tivit atstheorie, Preuss. Akad. Wiss. Berlin, Sitzber. 142
[5] Hartle, J.B.: 2002, Gravity: An Introduction to Einsteins General Relativity,
AddisonWesley
[6] Robertson, H.P.: 1935, Kinematics and world structure, ApJ 82, 248
[7] Straummann, N.: 2004, General Relativity. With Application to Astrophysics,
SpringerVerlag, Heidelberg (this is a more mathematically oriented approach to
General Relativity, with a complete introduction into modern differential geometry)
[8] Tammann, G., Parodi, B.R, Reindl, R.: 1999, in Proc. IAU Coll. 167, ASP Conf.
Series
[9] Walker, A.G.: 1936, On Milnes theory of worldstructure, Proc. Lond. Math. Soc.
42, 90
4 The Universe with Matter and Dark Energy
The FRW world models essentially only depend on the expansion factor R(t). The time
evolution for the expansion factor results from Einsteins equations, which couple curvature
with the matter content of the Universe. As a rst input we need a model for the description
of matter in the Universe, which includes all types of particles and uids.
4.1 Description of Matter in a Relativistic Cosmos
The present Universe contains various forms of matter: photons, neutrinos, baryons, galaxies,
dark matter and dark energy.
4.1.1 Fluid Approach
A great many astrophysical systems may be approximately regarded as perfect uids. A per-
fect uid is dened as having at each point a velocity v, such that an observer moving with
this velocity sees the uid around him as isotropic. This will be the case if the mean free path
between collisions is small compared with the scales used by the observer.
First we suppose that we are in a frame of reference in which the uid is at rest at some par-
ticular position and time. The perfect uid hypothesis tells us then that the energymomentum
tensor takes the form

T
00
= (4.1)

T
0i
= 0 (4.2)

T
ik
= P
ik
. (4.3)
The coefcients and P are proper energy density and pressure, respectively. Now we go
to a reference system at rest in the lab, and suppose that the uid in this frame appears to be
moving with velocity v. The connection between the two systems is given by a Lorentz boost

(v) with Lorentz factor = 1/


_
1

2
, where

= v/c. Since T is a symmetric tensor
of second rank, we have
T

(v)

(v)

T

. (4.4)
This gives explicitly
T
00
=
2
( +Pv
2
) (4.5)
T
i0
=
2
( +P)v
i
(4.6)
T
ik
= P
ik
+
2
( +P)v
i
v
k
. (4.7)
88 4 The Universe with Matter and Dark Energy
This can be integrated into a single equation for the energymomentum tensor of a perfect
uid
T

= ( +P) U

P g

, (4.8)
where U

is the velocity 4vector


U
0
=
dt
d
= ,

U =
dx
d
= v . (4.9)
Apart from energy and momentum, a uid will aslo carry a conserved number current.
If n is the particle number density in a Lorentz frame that moves with the uid, then in this
frame the particle current is given by

N
0
= n ,

N
i
= 0 . (4.10)
In any other Lorentz frame the expression follows from a boost transformation
N
0
=
0

(v)

N

= n (4.11)
N
i
=
i

(v)

N

= nv
i
, (4.12)
or more concisely
N

= nU

. (4.13)
This current is conserved,

= 0, and also the energymomentum is conserved, T

,
= 0.
This latter equations correspond to Euler equations

t
v + (v )v =
1 v
2
+P
(P +v
t
P) (4.14)
and the entropy conservation

t
s + (v )s = 0 (4.15)
for perfect uids. s denotes the specic entropy of the uid.
According to the covariance principle, the form of the energymomentum tensor is main-
tained when going to curved spaces with the same interpretation for and P, keeping the form
of Eq (4.8), and the equations of motion follow from the covariant divergence
T

;
= 0 , (4.16)
where the semicolon refers to the covariant derivative on the spacetime. Similarly, particle
number conservation follows from N

;
= 0.
4.1.2 Kinetic Description
Galaxies (pressureless uids, P = 0) and baryons can be modeled in terms of the uid ap-
proach. But photons, dark matter and neutrinos are collisionless species and require a kinetic
approximation. To describe a nonuniform system of particles in Minkowski spacetime one
introduces a local density n(t, x). This quantity is dened in such a way that n(t, x)
3
x
4.1 Description of Matter in a Relativistic Cosmos 89
gives the average number of particles in the spatial volume element
3
x at the position x and
time t. Similarly one denes a particle ow

j(t, x). Together they form a fourvector eld


N

= (cn(t, x),

j(t, x)). Let us consider a simple system of relativistic particles of mass m


with momenta p and energy cp
0
. If the number of particles is large, we introduce a function
f(x, p) which gives the distribution of 4momenta p

at each event. The denition is such


that f(x, p)
3
x
3
p gives the average number of particles which at time t are located in the
volume element
3
x at the point x with momenta lying in the range [ p, p + p].
Particle Current
With the help of this distribution function we dene the particle density as
n(t, x) =
_
d
3
p f(x, p) . (4.17)
In the same way we introduce the particle ow as

j(t, x) =
_
d
3
pv f(x, p) = n < v >, (4.18)
where v = c p/p
0
is the velocity of the relativistic particle with momentum p. The particle
owvector has therefore the covariant form
N

(x) = c
_
d
3
p
p
0
p

f(x, p) . (4.19)
Since the volume d
3
p/p
0
on the massshell is invariant under Lorentz transformations, the
phase space distribution f(x, p) must be a scalar object, d
3
p/p
0
is the Lorentzinvariant vol-
ume on the massshell.
EnergyMomentum Tensor
Since the energy per particle is cp
0
, the average energy density can be written as
T
00
= c
_
d
3
p p
0
f(x, p) , (4.20)
where T
00
denotes the energy density, as indicated by the above expression for perfect uids.
In a similar manner, we may dene the energy ow cT
0i
with
T
0i
=
_
d
3
p p
0
v
i
f(x, p) . (4.21)
Finally, we introduce the momentum ow (or pressure tensor) which is the ow in direction k
of the momentum in direction i
T
ik
=
_
d
3
p p
i
v
k
f(x, p) . (4.22)
In fact, this object is a twotensor which can be written in compact and covariant form
T

(x) = c
_
d
3
p
p
0
p

f(x, p) . (4.23)
90 4 The Universe with Matter and Dark Energy
Hence, in the relativistic kinetic theory, the energymomentum tensor is dened as the second
moment of the distribution function, and thus a symmetric tensor. The particle current is the
rst moment.
In this form, the energymomentum tensor takes only the rest energy and the kinetic en-
ergy of the particles into account. This is the case for dilute systems in the sense that the
interaction energy of the particles is small as compared to their kinetic energies. Otherwise,
the energymomentum tensor should include a potential energy contribution.
Entropy Flow
The Hfunction, introduced by Boltzmann, implies that the local entropy density for a system
outside equilibrium may be dened as
S
0
(x)/c = k
B
_
d
3
p f(x, p)
_
log[h
3
f(x, p)] 1
_
(4.24)
with a constant h such that h
3
f(x, p) is dimensionless. Conveniently, h is taken as the Planck
constant. The entropy ow has then the form

S(x) = k
B
_
d
3
pvf(x, p) [log[h
3
f(x, p)] 1] . (4.25)
This can be combined into a covariant form
S

(x) = k
B
c
_
d
3
p
p
0
p

f(x, p)
_
log h
3
f(x, p) 1
_
. (4.26)
Equilibrium Distributions
In the early universe, the plasma is extremely hot, completely ionized and so dense that typ-
ically interactions occur on timescales much shorter than the expansion timescale. Dense
plasmas are described in terms of a phasespace distribution f
eq
(t, p), also called occupation
number, which is isotropic in phasespace. As shown above, the number density n, the energy
density and the pressure P follow then from integrals over phasespace, given in the rest
system of the plasma,
n =
g
(2)
3
_
f
eq
(t, p) d
3
p (4.27)
c
2
=
g
(2)
3
_
E( p) f
eq
(t, p) d
3
p (4.28)
P =
g
(2)
3
_
| p|
2
3E
f
eq
(t, p) d
3
p . (4.29)
E is the relativistic energy of a particle corresponding to momentum p, and g is the spin
degeneracy factor for a particle species. For particles in kinetic equilibrium, the distribution
function is either the Fermi or Bosedistribution, which only depend on energy and temper-
ature, but not on angles in phasespace
f
eq
(| p|) =
1
exp[(E )/kT] 1
(4.30)
4.1 Description of Matter in a Relativistic Cosmos 91
with chemical potential .
In thermal equilibrium, these plasma parameters are then given by
n =
g
2
2

3
c
3
_

m

E
2
m
2
c
4
exp[(E )/kT] 1
E dE (4.31)
c
2
=
g
2
2

3
c
3
_

m

E
2
m
2
c
4
exp[(E )/kT] 1
E
2
dE (4.32)
P =
g
6
2

3
c
3
_

m
(E
2
m
2
c
4
)
3/2
exp[(E )/kT] 1
dE . (4.33)
These integrals can be evaluated for relativistic bosons (kT mc
2
, kT )
1
n
B
=
(3)

3
c
3
g (kT)
3
(4.35)

B
c
2
=

2
30
3
c
3
g (kT)
4
(4.36)
P
B
=
1
3

B
c
2
, (4.37)
and for relativistic fermions
n
F
=
3
4
(3)

3
c
3
g (kT)
3
=
3
4
n
B
(4.38)

F
c
2
=
7
8

2
30
3
c
3
g (kT)
4
=
7
8

B
c
2
(4.39)
P
F
=
1
3

F
c
2
. (4.40)
With respect to bosons (photons e.g.), the number density of fermions has a weight factor of
3/4 and the energy density a weight factor of 7/8.
Relativistic particles always satisfy the EOS P = c
2
/3.
4.1.3 EOS for Vacuum Energy
When longrange elds are present in the Universe, they also contribute to the energymomentum
tensor. As an example we consider a scalar eld (t, x) given by its Lagrangian density
L

=
1
2

V () (4.41)
with potential energy V (). This provides us the energymomentum tensor
T

_
1
2

V ()
_
. (4.42)
1
The Riemann function is dened as
(x) =
1
(x)
_

0
u
x1
exp(u) 1
du (4.34)
with its values (3) = 1.202... and (4) =
4
/90.
92 4 The Universe with Matter and Dark Energy
The corresponding density and pressure in a FRWmodel are given by

=
1
2

2
+
1
2
R
2
(t)()
2
+V () (4.43)
P

=
1
2

1
6
R
2
(t)()
2
V () . (4.44)
If there is a nonvanishing vacuum expectation value < >=
0
= 0, the energy density

= V (
0
) is constant and the corresponding pressure P

is negative for positive


V (
0
).
Any form of matter with an EoS of the form P < /3 is called Dark Energy.
4.2 Einsteins Equations for FRW Models
According to Einsteins equations, the total energymomentum tensor Tincluding all sorts of
matter is the source of gravity in the Universe
R
ab

1
2
g
ab
Rg
ab
= T
ab
. (4.45)
is the corresponding coupling constant determined by its Newtonian limit
=
8G
c
4
. (4.46)
Ris the Ricci scalar.
4.2.1 Derivation of Friedmanns Equations
We work in the conformal 3metric (i, k = 1, 2, 3)
ds
2
= dt
2

R
2
(t)
(1 +k
2
/4)
2

ik
dx
i
dx
k
(4.47)
with the corresponding natural oneforms

0
= dt (4.48)

i
=
R(t)
1 +k
2
/4
dx
i
,
2
=
3

i=1
(x
i
)
2
. (4.49)
In these forms, the metric is expressed as g =
a

b

ab
. It is easy to work out the
Christoffel symbols and from there the Riemann tensor in a coordinate basis. Here, I show
how the method with the Cartan equations can be used (Sect. 3.2.4). In this method, we have
rst to calculate the exterior derivatives of the observer frames (see Sect. 3.2.3)
d
0
= 0 (4.50)
d
i
=

R
R

0

i

k
2R
x
j

i

j
. (4.51)
With Cartans rst equation
d
a
=
a
b

b
(4.52)
4.2 Einsteins Equations for FRW Models 93
we can then solve for the connection forms

i0
=
0i
=

R
R

i
(4.53)

ij
=
ji
=
k
2R
[x
i

j
x
j

i
] , (4.54)
since the connection form has to be antisymmetric
dg

= 0 . (4.55)

i
j
is the connection of the slices t = const. With this we get the exterior derivatives of the
connection oneforms,
d
i0
=

R
R

0

i

k

R
2R
2
x
j

i

j
(4.56)
d
ij
=
k
R
2
_
1 +
k
2
4
_

i

j

k
2
4R
2
[x
i
x
m

j

m
x
j
x
m

i

m
] . (4.57)
This can now be used in the second structure equation

a
b
= d
a
b
+
a
c

c
b
, (4.58)
or for the individual components

0
i
= d
0
i
+
0
m

m
i
(4.59)
and

i
j
= d
i
j
+
i
0

0
j
+
i
m

m
j
. (4.60)
Inserting the above expressions, we can read off from these equations the curvature 2forms

0
i
=

R
R

0

i
(4.61)

i
j
=
k +

R
2
R
2

i

j
. (4.62)
This shows explicitly that the Riemann tensor is isotropic
R
0
i0k
=

R
R

ik
, R
0
ijm
= 0 (4.63)
R
i
kik
=
k +

R
2
R
2
, R
i
k0m
= 0 . (4.64)
The FRWmodel is a very simple spacetime where only two components of the Riemann
tensor are independent ! These expressions also demonstrate that the Riemann tensor does not
depend on the chosen coordinate system, since it is constant on 3surfaces.
94 4 The Universe with Matter and Dark Energy
The Ricci tensors can be derived from the above expressions
R
00
= R
1
010
+R
2
020
+R
3
030
= 3

R
R
(4.65)
R
11
= R
0
101
+R
2
121
+R
3
131
=
_

R
R
+ 2
k +

R
2
R
2
_
(4.66)
R
22
= R
11
(4.67)
R
33
= R
11
. (4.68)
This can be written in compressed form, now including the speed of ligth
2
R
00
=
3
c
2

R
R
(4.69)
R
0i
= 0 (4.70)
R
ik
=
1
c
2
_

R
R
+ 2
kc
2
+

R
2
R
2
_

ik
, (4.71)
as well as for the Ricciscalar
R =
6
c
2
_

R
R
+
kc
2
+

R
2
R
2
_
. (4.72)
With this, we obtain the Einstein tensor
G
00
=
3
c
2
kc
2
+

R
2
R
2
(4.73)
G
0i
= 0 (4.74)
G
ik
=
1
c
2
_
2

R
R
+
kc
2
+

R
2
R
2
_

ik
. (4.75)
The Einstein tensor together with the energymomentum tensor determines the Friedmann
equations

R
2
+kc
2
R
2

c
2

3
=
8G
3c
2
T
0
0
(4.76)
2

R
R
+

R
2
+kc
2
R
2
c
2
=
8G
c
2
T
1
1
. (4.77)
Here, we use the concrete expressions for the energymomentum tensor,
T
0
0
= c
2
, (4.78)
T
1
1
= T
2
2
= T
3
3
= P , (4.79)
resulting then in the Friedmann equations [6]

R
2
+kc
2
R
2
=
8G
3
+
c
2

3
(4.80)
2
Remember that these components are expressed in terms of orthonormal systems.
4.2 Einsteins Equations for FRW Models 95
and
2

R
R
+

R
2
+kc
2
R
2
=
8GP
c
2
+c
2
. (4.81)
By means of the rst equation, we may simplify the second one

R
R
=
4G(c
2
+ 3P)
3c
2
+
c
2

3
. (4.82)
The second Friedmann equation, which is known as the Raychaudhuri equation for the
tidal force R
0
i0i
=

R/R, implies the essential condition

R < 0, provided = 0, and

R > 0,
whenever vacuum energy is dominant.
Models including a dominant nonvanishing vacuum energy are always accelerated at
late times.
4.2.2 Energy Conservation
The expansion factor R(t) satises the following identity
d
dt
[R(

R
2
+kc
2
)] =

R[2R

R +

R
2
+kc
2
] (4.83)
and therefore as a consequence of the Friedmann equation,
d
dR
(c
2
R
3
) + 3P R
2
= 0 . (4.84)
This is known as the cosmic energy conservation in the form of the rst law of thermody-
namics for the energy E = c
2
R
3
dE +P dV = 0 , (4.85)
since V = R
3
is a measure for the comoving volume. For pressureless matter (galaxies or
dark matter), this is nothing else than the number conservation
(t) =
0
_
R
0
R(t)
_
3
. (4.86)
For relativistic matter, such as photons, neutrinos etc., with EOS P = c
2
/3 we nd
(t) =
0
_
R
0
R(t)
_
4
. (4.87)
For a general EOS of the form P = wc
2
with w = const the energy conservation leads
to the density evolution
(t) =
0
_
R
0
R(t)
_
3(1+w)
. (4.88)
This includes the above special cases. In particular for w = 1 the density must stay constant.
96 4 The Universe with Matter and Dark Energy
4.2.3 Density Parameters for Friedman Models
With the Hubble constant we can dene a characteristic or critical density

c

3H
2
0
8G
= 1.88 10
29
h
2
g cm
3
= 2.77 10
11
h
2
M

Mpc
3
. (4.89)
In modern Cosmology, it is now a tradition to parametrize the source terms in the Friedmann
equation by means of four parameters
H
2
(z) = H
2
0
_

R
(1 +z)
4
+
M
(1 +z)
3
+
k
(1 +z)
2
+

_
. (4.90)
The various parameters follow from the denition of the critical density
Density parameter for nonrelativistic matter:

M

8G
3H
2
0

M,0
=

M,0

c
(4.91)
Vacuum energy parameter:

=
c
2
3H
2
0
(4.92)
Curvature parameter:

k
=
kc
2
R
2
0
H
2
0
=
kR
2
H
R
2
0
. (4.93)
Curvature is small in a Universe, whenever the Hubble radius R
H
is much smaller than
the scaling radius R
0
. This is the basic idea behind inationary models. The observable
Universe is then practically at, which does not mean that it is globally at.
Radiation density parameter:

R

8G
3H
2
0

Rad,0
=

Rad,0

c
(4.94)
The radiation density consists of the CMB and the neutrino background, if neutrinos
are massless. The present contribution from CMB is negligibly small,

= (2.471
0.004) 10
5
/h
2
= (4.9 0.5) 10
5
.
These parameters satisfy, as a consequence of the Friedmann equation, the following relation

M
+

+
k
= 1 . (4.95)
Friedmann models of the cosmos are given in the cubus (
M
,

,
k
) by a plane, also called
the fundamental plane of cosmology. Baryons contribute a small fraction to the cosmic
matter

B
= (0.0224 0.0009)/h
2
= 0.044 0.004 . (4.96)
For this we can always write

M
=
B
/f
Gas
(4.97)
with 0.1 f
Gas
0.2 from observations in galaxy clusters. This gives a present value of

M
= (0.135 0.08)/h
2
= 0.27 0.04 . (4.98)
4.3 FRW Models without Vacuum Energy 97
4.3 FRW Models without Vacuum Energy
At present time, the Universe is dominated by cold matter, i.e. matter in galaxies and dark
matter. Also, the energy density in the CMBR is many orders of magnitude below the critical
density and has not to be included in the source terms for gravity. Hence, neglecting vacuum
energy, the above equations simplify

R
2
+kc
2
R
2
=
8G
3
(4.99)
2

R
R
+

R
2
+kc
2
R
2
= 0 . (4.100)
The rst equation now determines the Hubble constant, t = t
0
,
H
2
0
+
kc
2
R
2
0
=
8G
0
3
. (4.101)
If the present matter density
0
were known exactly, we could decide whether the Universe
is presently closed or open, since
M
+
k
= 1. Compared to present densities, the critical
density seems to dominate all forms of matter. The contribution for galaxies to the density
parameter
M
is small

G
0.005 . (4.102)
The Universe could however be dominated by some form of dark matter (invisible to tele-
scopes), such that
M
1. The density parameter
M
is therefore the second parameter
which determines the dynamical state of the present Universe. Models with

= 0 and

M
= 1 were called Standard Cold Dark Matter models, or SCDM. These models were
the standard model for Cosmology in the 90s.
As we have seen, it is also common to describe the dynamical state in terms of a deceler-
ation parameter
q
0
H
2
0
=
_

R
R
_
0
, (4.103)
which is related to the curvature of the expansion factor a(t). For models with vanishing
cosmological constant we nd the important relation
q
0
=
1
2

M
,

= 0 . (4.104)
According to this, the Universe is at, if
M
= 1 (q
0
= 0.5), closed for
M
> 1 and open
for
M
< 1.
The energy conservation provides us the evolution of the density as a function of the
present density
0
(t) =
0
_
R
0
R(t)
_
3
=
M

c
_
R
0
R(t)
_
3
. (4.105)
With this expression we can determine the solution of the Friedmann equation for all values
of k.
98 4 The Universe with Matter and Dark Energy
4.3.1 The Euclidean Universe, k=0
For k = 0 we nd the equation

R
2
=
8G
0
3
R
3
0
R
= H
2
0
R
3
0
R
, (4.106)
which has the simple solution, called Einsteinde Sitter model,
R(t) = R
0
_
t
t
0
_
2/3
. (4.107)
The Hubble constant gives then the age of the Universe
t
0
=
2
3H
0
. (4.108)
This also yields time as a function of redshift
t(z) = t
0
(1 +z)
3/2
. (4.109)
4.3.2 The Closed Universe, k=1
For k = 1 the Friedman equation yields
2

R
R
+

R
2
+c
2
R
2
= 0 (4.110)

R
2
+c
2
R
2

8G
0
R
3
0
3R
3
= 0 . (4.111)
The second equation is the essential equation which gives
_

R
R
0
_
2
= H
2
0
_
1 2q
0
+ 2q
0
R
0
R
_
. (4.112)
Its solution is given as t = t(R)
t =
1
H
0
_
R/R
0
0
_
1 2q
0
+
2q
0
x
_
1/2
dx (4.113)
with the present age dened as
t
0
=
1
H
0
_
1
0
_
1 2q
0
+
2q
0
x
_
1/2
dx <
1
H
0
. (4.114)
In this case one obtains the solution of the Friedman equation over the transformation
1 cos =
2q
0
1
q
0
R(t)
R
0
. (4.115)
4.3 FRW Models without Vacuum Energy 99
The relation between and time follows from the equation (4.112)
H
0
t =
q
0
(2q
0
1)
3/2
(sin ) . (4.116)
This represents now an implicit equation for the parameter as a function of time. Allto-
gether, this represents the equation for a cycloid with its maximum at
m
= , i.e. at the
time
t
m
=
q
0
H
0
(2q
0
1)
3/2
, R
m
=
2q
0
R
0
2q
0
1
. (4.117)
For the present time we nd therefore
0
= (t
0
), or
cos
0
=
1 q
0
q
0
, sin
0
=

2q
0
1
q
0
. (4.118)
With this information we are able to calculate the age of the closed Universe
t
0
=
R
m
2c
(
0
sin
0
) (4.119)
=
1
H
0
q
0
(2q
0
1)
3/2
_
cos
1
_
1 q
0
q
0
_

2q
0
1
q
0
_
. (4.120)
As example, one obtains for q
0
= 1
t
0
=
1
H
0
_

2
1
_
. (4.121)
In comparison to a at Universe, the closed Universe is younger.
The solution for the expansion factor describes a cycloid (see Fig. 4.1). The radius of the
Universe reaches its maximum at = with the value
R
max
= R
m
=
c
H
0
2q
0
(2q
0
1)
3/2
. (4.122)
A closed Universe cycles therefore between expansion and contraction and reaches its mini-
mum R = 0 after a total time
t
L
=
R
m
c
=
1
H
0
2q
0
(2q
0
1)
3/2
. (4.123)
For q
0
= 1, the lifecycle of the Universe is t
L
= 2/H
0
.
4.3.3 The Open Universe, k=1
Finally, we consider the case k = 1 with the Friedman equation
2

R
R
+

R
2
c
2
R
2
= 0 (4.124)

R
2
c
2
R
2

8G
0
R
3
0
3R
3
= 0 . (4.125)
100 4 The Universe with Matter and Dark Energy
The essential dynamics is hidden in the equation

R
2
= c
2
_
1 +
R
m
R
_
(4.126)
with
R
m
=
2q
0
(1 2q
0
)
3/2
c
H
0
. (4.127)
Similar to the closed Universe, one can obtain the solution in terms of the transformation
R(t) =
R
m
2
(cosh (t) 1) , ct =
R
m
2
(sinh ) . (4.128)
This is valid in the range 0 q
0
< 1/2, or 0 < 1. The present value of is given by
cosh
0
=
1 q
0
q
0
, sinh
0
=

1 2q
0
q
0
. (4.129)
This gives the age of the Universe in terms of the Hubble age t
H
= 1/H
0
t
0
=
R
m
2c
(sinh
0

0
) (4.130)
=
1
H
0
q
0
(1 2q
0
)
3/2
_
1 2q
0
q
0
ln
_
1 q
0
+

1 2q
0
q
0
_
_
. (4.131)
4.3.4 The Singularity at t=0: BigBang
All three Friedman models have the genuine property that R = 0 is reached after a nite time.
Near R = 0, the Hubble parameter explodes and becomes nally unbounded. At the same
time, we observe that all components of the Riemann tensor become singular at this point.
This point is therefore called BigBang.
This singularity is an inherent property of the Einstein equations. The curvature of the the
solution R(t) is always negative

R < 0, falls c
2
+3P > 0. This energy condition is certainly
satised for all type of classical matter, even in the early Universe at high densities. If this
energy condition is never violated, the Universe has to pass through a singular state at t = 0.
One can even show that the singularity survives in even less symmetric spaces. At such small
scales, however, quantum effects are expect to play an essential role (space will be quantised).
4.3.5 The Mattig Formula for the Luminosity Distance
As an example of the general considerations in Sect. 3.7 we want to calculate the luminosity
distance for these classical Friedman models, d
L
= r
em
R(t
0
)(1 +z).
4.3 FRW Models without Vacuum Energy 101
Figure 4.1: The expansion factor of classical Friedman models.
Flat Universe
In the at case we simply get
r
1
=
_
t
0
t
1
c dt
R(t)
=
c
R
0
_
t
0
t
1
t
2/3
0
t
2/3
dt
= 3
c
R
0
t
2/3
0
(t
1/3
0
t
1/3
1
)
=
3c
R
0
t
0
_
1
_
t
1
t
0
_
1/3
_
. (4.132)
Together with redshift this gives
r
1
=
3ct
0
R
0
[1 (1 +z)
1/2
] =
2c
R
0
H
0
[1 (1 +z)
1/2
] . (4.133)
The luminosity distance follows therefore from
d
L
= r
1
R
0
(1 +z) =
2c
H
0
[(1 +z)

1 +z] . (4.134)
This correspond to the classical Hubblelaw for z 1, d
L
(c/H
0
)z.
102 4 The Universe with Matter and Dark Energy
Closed Universe
In the closed Universe we nd
_
r
1
0
dr

1 r
2
=
_
t
0
t
1
c dt
R(t)
. (4.135)
The left integral is simple, sin
1
r
1
, and the right hand side follows from
_
t
0
t
1
c dt
R(t)
=
_
t
0
t
1
dR
_
R(R
m
R)
=
_

0

1
d =
0

1
. (4.136)
Therefore
r
1
= sin(
0

1
) , (4.137)
and together with redshift we have
1 +z =
R(t
0
)
R(t
1
)
=
sin
2
(
0
/2)
sin
2
(
1
/2)
(4.138)
the relation
sin
1
=
2
1 +z
sin

0
2
_
z + cos
2

0
2
(4.139)
cos
1
=
z + cos
0
1 +z
. (4.140)
On the other hand, the parameter
0
satises the Friedman equation
sin

0
2
=
_
2q
0
1
2q
0
(4.141)
cos

0
2
=
_
1
2q
0
. (4.142)
With this we obtain the expression for r
1
r
1
=

2q
0
1
q
2
0
(1 +z)
_
q
0
z + (1 q
0
)[1
_
1 + 2q
0
z]
_
. (4.143)
This provides the wellknown relation for the distance as a function of redshift, the socalled
Mattig formula
d
L
= r
1
R
0
(1 +z) =
c
H
0
1
q
2
0
_
q
0
z + (q
0
1)[
_
1 + 2q
0
z 1]
_
. (4.144)
Be aware of the fact that for z 1 the parameter q
0
disappears, and we nd once again the
classical Hubblelaw. As already discussed in Sect. 3.7, the Hubblelaw is independent of
the curvature of the space.
4.4 The Present Universe with Dark Energy 103
Open Universe
For open Universe, the calculation is similar
r
1
=

1 2q
0
q
2
0
(1 +z)
_
q
0
z + (1 q
0
)[1
_
1 + 2q
0
z]
_
, (4.145)
and therefore for the distance
d
L
=
c
H
0
1
q
2
0
_
q
0
z + (q
0
1)[
_
1 + 2q
0
z 1]
_
. (4.146)
This expression is exactly the expression we found for the closed Universe, and in the limit
Dieser Ausdruck unterscheidet sich nicht von dem im geschlossenen Modell, und q
0
1/2
we also get the luminosity distance of the at Universe Therefore, equation (4.144) is valid
for all three types of models, i.e. for all values of q
0
0.
4.4 The Present Universe with Dark Energy
Results from WMAP and the analysis of the luminosity distance for high redshift Supernovae
indicate that the present state of the Universe cannot be satisfactorily described in terms of
classical Friedmann models not including vacuum energy. We will discuss in the following
soltuions of the Friedmann equation including some form of vacuum energy given by an EOS
of the form (see the discussion in 3.2)
P
V
= w
V

V
. (4.147)
As we have seen in 3.2, a cosmological constant is equivalent to a vacuum energy density
dened as

=
c
2
8G
. (4.148)
The total energy density of the present Universe has then three contributions
(t) =
M
(t) +
Rad
(t) +
V
, (4.149)
and similarly for the pressure
P(t) = P
M
(t) +P
Rad
(t) +P
V
, P
V
= w
V
c
2
(4.150)
with w
V
= 1.
The energy density
V
of the vacuum always results from ground state energy in eld
theories. Both terms
V
and

occur as sum in Einsteins equations auf


_

R
R
_
2
=
8G
3
(
M
+
Rad
+
V
+

)
kc
2
R
2
, (4.151)

R
R
=
4G
3
(
M
+ 6P
Rad
2
V
c
2
2

c
2
) . (4.152)
104 4 The Universe with Matter and Dark Energy
The sum of all energy densities cannot excede the present critical density, i.e. 4 10
56
cm
2
. Natural values for the vacuum energy density are either given by QCD or the Planck
epoch

QCD
V
c
2
= (0.3 GeV )
4
= 1.6 10
36
erg/cm
3
(4.153)

Planck
V
c
2
= (10
18
GeV )
4
= 2 10
110
erg/cm
3
. (4.154)
It is one of the big mysteries, why this vacuum energy is so small,
V

crit
.
4.4.1 Cosmological Parameters
It is usual to normalize the radius to its present value, a(t) = R(t)/R
0
. Then the Friedmann
equation simplies to
a
2
= H
2
0
(

M
a
+
V
a
2
)
kc
2
R
2
0
. (4.155)
Time is measured in units of H
1
0
.
V
is the density parameter of the vacuum.
For t = t
0
this is equivalent to
1 =
M
+
V

kc
2
R
2
0
H
2
0
, (4.156)
or to
R
0
=
c
H
0
_
k

M
+
V
1
. (4.157)
From here, we get the differential equation
a
2
= H
2
0
_

M
a
+
V
a
2
+ 1
M

V
_
. (4.158)
This corresponds to the equation of motion of a particle in a potential
M
/x
V
x
2
with
energy 1
M

V
. For
V
> 0 the potential has a maximum and it is negative for all
values x, i.e. the solutions with k = 0, 1 expand away from a singularity. For k = +1 we
nd a critical value
V
, so that

R
2
= 0.
As in classical Friedmann models, the deceleration parameter q
0
is quite often used to
parametrize the present state of the Universe
q
0
=

RR

R
2
=

R
RH
2
. (4.159)
Together with the second Friedmannequation,

R/R = 4GR( + 3P/c
2
)/3 +c
2
/3 we
obtain the relation
q
0
=
1
2

M

. (4.160)
q
0
is therefore not an independent parameter.
The existence of the vacuum energy determines therefore the geometry of the Uni-
verse (Fig. 4.3). A nonvanishing vacuum energy leads to a at Universe, even if the
matter density is smaller than the critical density. A positive vacuum energy density
accelerates the expansion of the Universe. In the follwoing, we only discuss models with
w = 1 (Fig. 4.3).
4.4 The Present Universe with Dark Energy 105
Figure 4.2: Expansion of the Universe dominated by vacuum energy. The transition from deceleration
to acceleration occurs around redshift one.
4.4.2 Solutions for Inationary Universes
The solutions of the Friedmann equation including vacuum energy are not trivial, but must
be determined by means of numerical techniques. There is however a special case, which is
analytically solvable. In a at Universe, the Friedmann equation is simply
_
a
a
_
2
= H
2
0
_

M
a
3
+ 1
M
_
. (4.161)
One can easily show that in the at case, k = 0, the following ansatz satises the above
Friedmann equation
a(t) =
_
_

M
1
M
sinh
_
3

1
M
H
0
t
2
_
_
2/3
(4.162)
As expected, for late times we nd an exponential expansion, a(t) exp(

1
M
H
0
t),
while the Taylor expansion for small timescales, sinh(x) x provides the classical expan-
sion law of a at Friedmann Universe with vanishing vacuum energy, a(t) t
2/3
(Fig. 4.4).
4.4.3 Age of the Universe
We write the Friedmann equation for the dimensionless expansion factor a(t) R(t)/R
0
=
1/(1 +z) with a(t
0
) = 1
da
dt
= H
0
_

M
a
+

a
2
+ 1
M

(4.163)
106 4 The Universe with Matter and Dark Energy
Figure 4.3: Transition from a decelerating expansion to the accelerating Universe dominated by dark
energy (top). The lower panel shows the future evolution of the Universe dominated by vacuum energy.
From this expression we can derive the age t = t(z) as function of redshift z, parametrized
by H
0
,
M
and

. Since
dt =
dt
da
da =
dt
da
dz
(1 +z)
2
=
1
H
0
dz
(1 +z)
2
_

M
(1 +z) +

/(1 +z)
2
+ 1
M

=
1
H
0
dz
(1 +z)
_

M
(1 +z)
3
+ (1
M

)(1 +z)
2
+

, (4.164)
4.4 The Present Universe with Dark Energy 107
Figure 4.4: Expansion of the Universe for vanishing curvature (inationary cosmos). The scale
factor is normalized to the present radius, time is measured in billions of years from today.
Curves in the blue region represent models with accelerated expansion, with a vacuum energy
of 95% to 40% of the critical density.
we nd for the age of the Universe as a function of redshift
t(z) =
_
t
0
dt =
1
H
0
_

z
dz

(1 +z

) E(z

)
(4.165)
with E(z) = H(z)/H
0
. This expression can easily be integrated numerically for any Fried-
man equation H = H(z) (see Fig. 4.6).
The special case of the at inationary Universe Eq (4.162),
k
= 0, can be simplied
to
t(z) =
_
t
0
dt =
1
H
0
_

z
dz

(1 +z

)
_

M
(1 +z

)
3
+

. (4.166)
With the transformation
tan (z) =
_

(1 +z)
3/2
(4.167)
108 4 The Universe with Matter and Dark Energy
Figure 4.5: Age of the Universe as a function of redshift z. The upper curve represents a at
CDMmodel with
M
= 0.3,

= 0.7, the lower curve (dashed line) a classical CDM


model (SCDM) with
M
= 1. Hubble constant H
0
= 65 km/s/Mpc.
the age can be solved analytically (Fig. 4.5)
t(z) =
2
3H
0

ln
_
1 + cos (z)
sin (z)
_
. (4.168)
The Empirical Age of the Universe
There are at least 3 ways that the age of the Universe can be estimated. The most important
ones are:
The age of the oldest star clusters (globular clusters).
The age of the chemical elements.
The age of the oldest white dwarf stars.
As we have seen, the age of the Universe can also be estimated from a cosmological model
based on the Hubble constant and the densities of matter and dark energy. This model-based
age is currently 13.7 0.2 Gyr. The actual age measurements are in fact consistent with the
modelbased age. This increases our condence in the Big Bang model.
Age of Globular Clusters: The life cycle of a star depends upon its mass. High mass
stars are much brighter than low mass stars, thus they rapidly burn through their supply of
hydrogen fuel. A star like the Sun has enough fuel in its core to burn at its current brightness
for approximately 11 billion years. A star that is twice as massive as the Sun will burn through
4.4 The Present Universe with Dark Energy 109
Figure 4.6: Age of the Universe as a function of
m
and

. The numbers of the contour lines


give the age of the Universe in terms of the Hubble age t
H
= 1/H
0
.
110 4 The Universe with Matter and Dark Energy
its fuel supply in only 800 million years. A 10 solar mass star, a star that is 10 times more
massive than the Sun, burns nearly a thousand times brighter and has only a 20 million year
fuel supply. Conversely, a star that is half as massive as the Sun burns slowly enough for its
fuel to last more than 20 billion years.
All of the stars in a globular cluster formed at roughly the same time, thus they can serve as
cosmic clocks. If a globular cluster is more than 20 million years old, then all of its hydrogen
burning stars will be less massive than 10 solar masses. This implies that no individual hydro-
gen burning star will be more than 1000 times brighter than the Sun. If a globular cluster is
more than 2 billion years old, then there will be no hydrogen-burning star more massive than
2 solar masses.
When stars are burning hydrogen to helium in their cores, they fall on a single curve in
the luminosity-temperature plot known as the HR diagram after its inventors, Hertzsprung and
Russell. This track is known as the main sequence, since most stars are found there. Since
the luminosity of a star varies like M
3
or M
4
, the lifetime of a star on the main sequence
varies like t =const M/L = k/L
0.7
. Thus if you measure the luminosity of the most lu-
minous star on the main sequence, you get an upper limit for the age of the cluster: Age
< k/L(MS
max
)
0.7
.
This is an upper limit because the absence of stars brighter than the observed L(MS
max
)
could be due to no stars being formed in the appropriate mass range. But for clusters with
thousands of members, such a gap in the mass function is very unlikely, the age is equal to
k/L(MS
max
)
0.7
. Chaboyer, Demarque, Kernan and Krauss (1996, Science 271, 957) apply
this technique to globular clusters and nd that the age of the Universe is greater than 12.07
Gyr with 95% condence. They say the age is proportional to one over the luminosity of the
RR Lyra stars which are used to determine the distances to globular clusters. Chaboyer (1997)
gives a best estimate of 14.6 1.7 Gyr for the age of the globular clusters. But Hipparcos
results have shown that the globular clusters are further away than previously thought, so their
stars are more luminous. Gratton et al. give ages between 8.5 and 13.3 Gyr with 12.1 being
most likely, while Reid gives ages between 11 and 13 Gyr, and Chaboyer et al. give 11.51.3
Gyr for the mean age of the oldest globular clusters.
The Age of the Elements: The age of the chemical elements can be estimated using radioac-
tive decay to determine how old a given mixture of atoms is. The most denite ages that can
be determined this way are ages since the solidication of rock samples. When a rock solidi-
es, the chemical elements often get separated into different crystalline grains in the rock. For
example, sodium and calcium are both common elements, but their chemical behaviours are
quite different, so one usually nds sodium and calcium in different grains in a differentiated
rock. Rubidium and strontium are heavier elements that behave chemically much like sodium
and calcium. Thus rubidium and strontium are usually found in different grains in a rock. But
Rb-87 decays into Sr-87 with a half-life of 47 billion years. And there is another isotope of
strontium, Sr-86, which is not produced by any rubidium decay. The isotope Sr-87 is called
radiogenic, because it can be produced by radioactive decay, while Sr-86 is non-radiogenic.
The Sr-86 is used to determine what fraction of the Sr-87 was produced by radioactive decay.
This is done by plotting the Sr-87/Sr-86 ratio versus the Rb-87/Sr-86 ratio. When a rock is
rst formed, the different grains have a wide range of Rb-87/Sr-86 ratios, but the Sr-87/Sr-86
ratio is the same in all grains because the chemical processes leading to differentiated grains
do not separate isotopes. After the rock has been solid for several billion years, a fraction of
the Rb-87 will have decayed into Sr-87. Then the Sr-87/Sr-86 ratio will be larger in grains
4.4 The Present Universe with Dark Energy 111
with a large Rb-87/Sr-86 ratio.
The Age of Oldest White Dwarfs: A white dwarf star is an object that is about as heavy as
the Sun but only the radius of the Earth. The average density of a white dwarf is a million times
denser than water. White dwarf stars form in the centers of red giant stars, but are not visible
until the envelope of the red giant is ejected into space. When this happens the ultraviolet
radiation from the very hot stellar core ionizes the gas and produces a planetary nebula. The
envelope of the star continues to move away fromthe central core, and eventually the planetary
nebula fades to invisibility, leaving just the very hot core which is now a white dwarf. White
dwarf stars glow just from residual heat. The oldest white dwarfs will be the coldest and thus
the faintest. By searching for faint white dwarfs, one can estimate the length of time the oldest
white dwarfs have been cooling. Oswalt, Smith, Wood and Hintzen (1996, Nature, 382, 692)
have done this and get an age of 9.5 + 1.1 0.8 Gyr for the disk of the Milky Way. They
estimate an age of the Universe which is at least 2 Gyr older than the disk, so t
0
> 11.5 Gyr.
Hansen et al. have used the HST to measure the ages of white dwarfs in the globular cluster
M4, obtaining 12.7 0.7 Gyr. In 2004 Hansen et al. updated their analysis to give an age for
M4 of 12.10.9 Gyr, which is very consistent with the age of globular clusters from the main
sequence turnoff. Allowing for the time between the Big Bang and the formation of globular
clusters (and its uncertainty) implies an age for the Universe of 12.8 1.1 Gyr.
4.4.4 The Event Horizon
We consider an event (t
1
, r
1
) which we wish to observe at our location r = 0
_
r
1
0
dr

1 kr
2
=
_
t
t
1
c dt

R(t

)
. (4.169)
An observer at r = 0 will be able to receive signals from any event (after a suitable long
wait), provided the integral on the rhs diverges (t ). Fro R(t) t
p
, this implies p < 0,
or a decelerating Universe. In an accelerating Universe, the integral converges, signaling the
presence of an event horizon. For this we need
_
r
1
0
dr

1 kr
2

_

0
c dt

R(t

)
. (4.170)
Observers beyond a distance
R
H
= R
0
_

t
0
c dt
R(t)
(4.171)
are not able to communicate with the rest of the world. For a de Sitter Universe R(t) =
R
1
exp(H(t t
1
)), this limit is R
H
= c/H.
4.4.5 The Particle Horizon
Alternatively, we may ask the question: Can we nowadays see the entire cosmos ? Or is there
a limit r
H
for r
1
, provided z ? This would imply the existence of R
H
satisfying
R
H
= R
0
_
t
0
t
e
0
c dt
R
= R
0
_
r
H
0
dr

1 kr
2
. (4.172)
112 4 The Universe with Matter and Dark Energy
In an Einsteinde Sitter cosmos with R(t) t
2/3
, t 0, this limit will be achieved
R
H
=
3c
H
0
, k = 0 . (4.173)
For a radiationdominated Universe, R(t)

t we nd R
H
= 2c/H. This is called particle
horizon: particles with r
1
> R
H
are hidden for us.
As an application we consider the particle horizon at recombination, where t
R
10
6
years. The particle horizon at that time was R
H
= 2c/H(t
R
). At that time, the domain of
inuence was much smaller than today, and despite this fact, we observe a large degree of
isotropy in the background radiation. This fact is called the causality problem of the standard
Universe (Fig. 4.7).
Figure 4.7: Causality problem at recombination: at redshift z = 1100 (blue line) we observe
radiation from patches of spaces which were not in causal contact. The time trajectories are
given here in conformal time, d = dt/R(t).
4.4.6 Conformal Maps of the Present Universe
Cartographers mapping the Earths surface were faced with the challenge of mapping a curved
surface onto a plane. No such projection can be perfect, but it can capture important features.
Perhaps the most famous map projection is the Mercator projection (presented by Gerhardus
Mercator in 1569). This is a conformal projection which preserves shapes locally. Lines
of latitude are shown as straight horizontal lines, while meridians of longitude are shown as
straight vertical lines.
The HammerAitoff projection shows the Earth as a horizontal ellipse with 2:1 axis ratio.
The equator is shown as a straight horizontal line marking the long axis of the ellipse. It is
produced in the following way: map the entire sphere onto its western hemisphere by simply
compressing each longitude by a factor of 2. Now map this western hemisphere onto a plane
by the Lambert equal area azimuthal projection. This map is a circular disk. This is then
stretched by a factor of 2 (undoing the previous compression by a factor of 2) in the equatorial
direction to make an ellipse with a 2:1 axis ratio. Thus, the Hammer projection preserves
areas.
4.4 The Present Universe with Dark Energy 113
In galaxy surveys we use slice maps of the universe to make at maps. The CfA maps
surveyed a slice of sky, 117 degrees long and 6 degrees wide, of constant declination. In 3D
this slice had the geometry of a cone, and they attened this onto a plane. (A cone has zero
Gaussian curvature and can therefore be constructed from a piece of paper. A cone cut along
a line and attened onto a plane looks like a pizza with a slice missing .) If the cone is at
declination , the map in the plane will be x = r cos(cos()), y = r sin(cos()), where
is the right ascension (in radians), and r is the co-moving distance (as indicated by the redshift
of the object). This will preserve shapes. Many times a 360 degree slice is shown as a circle
with the Earth in the center, where x = r cos(), y = r sin(). If r is measured in co-moving
distance, this will preserve shapes only if the universe is at (k = 0), and the slice is in the
equatorial plane ( = 0), (if = 0, structures (such as voids) will appear lengthened in the
direction tangential to the line of sight by a factor of 1/ cos()). Therefore, it is important
to investigate map projections which will preserve shapes locally. If one has the correct
cosmological model, and uses such a conformal map projection, isotropic features in the large
scale structure will appear isotropic on the map.
Our objective here is to produce a conformal map of the universe which will show the
wide range of scales encountered while still showing shapes that are locally correct. Consider
the general Friedmann metrics in the spherical coordinates
ds
2
= dt
2
+a
2
(t)(d
2
+ sin
2
(d
2
+ sin
2
d
2
)), k = +1 (4.174)
ds
2
= dt
2
+a
2
(t)(d
2
+
2
(d
2
+ sin
2
d
2
)), k = 0 (4.175)
ds
2
= dt
2
+a
2
(t)(d
2
+ sinh
2
(d
2
+ sin
2
d
2
)), k = 1 (4.176)
where t is the cosmic time since the Big Bang, a(t) is the expansion parameter, and individual
galaxies participating in the cosmic expansion follow geodesics with constant values of , ,
and . These three are called co-moving coordinates. Neglecting peculiar velocities, galaxies
remain at constant positions in co-moving coordinates as the universe expands. The expansion
factor a(t) obeys Friedmanns equations.
We can dene a conformal time by the relation d = dt/a, so that
(t) =
_
t
0
dt
a
(4.177)
Light travels on radial geodesics with d = d so a galaxy at a co-moving distance from
us emitted the light we see today at a conformal time (t) = (t
0
). Thus, we can calculate
the time t and redshift z = a(t
0
)/a(t) 1 at which that light was emitted. Conversely, if we
know the redshift, given a cosmological model we can calculate the co-moving radial distance
of the galaxy from us from its redshift, again ignoring peculiar velocities.
The WMAP data implies that w 1 for dark energy (ie. p
vac
= w
vac

vac
),
suggesting that a cosmological constant is an excellent model for the dark energy, so we are
simply adopting that. The current Hubble radius R
H
0
= c/H
0
= 4220 Mpc. The cosmic
microwave background is at a redshift z = 1089. Substituting, using geometrized units in
which c = 1, and integrating the rst Friedmann equation we nd the conformal time may be
calculated
(t) =
_
t
0
dt
a
=
_
a(t)
0
da
a
2
H(a)
=
1
H
0
_
a(t)
0
_

r
+
m
a +
k
a
2
+

a
4
_
1/2
da
(4.178)
114 4 The Universe with Matter and Dark Energy
where
m
a
3
and
r
a
4
. This formula will accurately track the value of (t), provid-
ing that this is interpreted as the value of the conformal time since the end of the inationary
period at the beginning of the universe. (During the inationary period at the beginning of
the universe, the cosmological constant assumed a large value, different from that observed
today, and the formula would have to be changed accordingly. So we simply start the clock
at the end of the inationary period where the energy density in the false vacuum [large cos-
mological constant] is dumped in the form of matter and radiation. Thus, when we trace back
to the big bang, we are really tracing back to the end of the inationary period. After that,
the model does behave just like a standard hot-Friedmann big bang model. This standard
model might be properly referred to as an inationary-big bang model, with the inationary
epoch producing the Big Bang explosion at the start.) Now, a(t) is the radius of curvature
of the universe for the k = +1 and k = 1 cases, but for the
k
= 0 case, which we will
be investigating rst and primarily, there is no scale and so we are free to normalize, set-
ting a(t
0
) = R
H
0
= c/H
0
= 4220 Mpc. Then, measures co-moving distances at the
present epoch in units of the current Hubble radius R
H
0
. Thus, for the
k
= 0 case, using
geometrized units, we have
(a) = (a(t)) =
_
a
0
_
a
a
0

m
+
r
+ (
a
a
0
)
4

_
1/2
da
a
0
(4.179)
where
m
,

,
r
are the values at the current epoch. Given the values adopted from WMAP
we nd
(a
0
) = 3.38 (4.180)
That means that when we look out now at t = t
0
(when a = a
0
) we can see out to a distance
of
= 3.38 (4.181)
or a co-moving distance of
R
H
0
= 3.38R
H
0
= 14, 300 Mpc . (4.182)
This is the effective particle horizon, where we are seeing particles at the moment of the
Big Bang. This is a larger radius than 13.7 billion light years the age of the universe (the
lookback time) times the speed of light because it shows the co-moving distance the most
distant particles we can observe now will have from us when they are as old as we are now, i.e.
measured at the current cosmological epoch. We may calculate the value of as a function of
a, or equivalently as a function of observed redshift z = (a
0
/a) 1. Recombination occurs
at z
rec
= 1089, which is the redshift of the cosmic microwave background seen by WMAP.
(z
rec
) = 0.0671 (4.183)
So, the co-moving radius of the cosmic microwave background is
R
H
0
= (
0

rec
)R
H
0
= 14, 000 Mpc (4.184)
That is the radius at the current epoch, so at recombination the WMAP sphere has a physical
radius that is 1090 times smaller or about 13 Mpc.
4.4 The Present Universe with Dark Energy 115
Redshift z r(z) (Mpc) Remark
14,283 Big Bang (end of inationary period)
3233 14,165 Equal matter and radiation density epoch
1089 14,000 Recombination
6 8,422
5 7,933
4 7,305
3 6,461
2 5,245
1 3,317
0.5 1,882
0.2 809
0.1 413
Table 4.1: Co-moving radii for different redshifts
We may compute co-moving radii r = R
H
0
for different redshifts, as shown in table 4.1.
We can also calculate the value of (t = ) = 4.50 which shows how far a photon can travel
in co-moving coordinates from the inationary Big Bang to the innite future. Thus, if we
wait until the innite future we will eventually be able to see out to a co-moving distance of
r
t=
= 4.50R
H
0
= 19, 000 Mpc . (4.185)
This is the co-moving future visibility limit, or future horizon. No matter how long we
wait, we will not be able to see further than this. This is surprisingly close. The number of
galaxies we will eventually ever be able to see is only larger than number observable today by
a factor of (r
t=
/r
t
0
)
3
= 2.36.
If we send out a light signal now, by t = it will reach a radius = (t = )(t
0
) =
4.50 3.38 = 1.12, or
r = 4, 740 Mpc (4.186)
to which we refer to as the outward limit of reachability. We cannot reach (with light signals
or rockets) any galaxies that are further away than this. What redshift does this correspond
to? Galaxies we observe today with a redshift of z = 1.69 are at this co-moving distance.
Galaxies with redshifts larger than 1.69 today are unreachable. This is a surprisingly small
redshift.
We can see many galaxies at redshifts larger than 1.69 that we will never be able to visit or
signal. In the accelerating universe, these galaxies are accelerating away from us so fast that
we can never catch them. The total number of stars that our radio signals will ever pass is of
order 2 10
21
.
116 4 The Universe with Matter and Dark Energy
A Map Projection for the Universe
We will choose a conformal map that will cover the wide range of scales from the Earths
neighborhood to the cosmic microwave background. First we will consider the at case (
k
=
0) which the WMAP data tells us is the appropriate cosmological model. Our map will be two
dimensional so that it can be shown on a wall chart. CfA survey showed with their slice of the
Universe, just how successful a slice of the Universe can be in illustrating large scale structure.
The Sloan Digital Survey includes spectra and accurate positions for about 1 million galaxies
and quasars in a 3D sample. We only use an equatorial slice 4 degrees wide (2
o
< < 2
o
)
centered on the celestial equator covering both northern and southern galactic hemispheres.
This shows many interesting features including many prominent voids and a great wall longer
than the great wall found in CfA.
Since the observed slice is already in a at plane (k = 0 model, along the celestial equa-
tor) we may project this slice directly onto a at sheet of paper using polar coordinates with
r = R
H
0
being the co-moving distance, and being the right ascension. We wish to show
large scale structure and the extent of the observable universe out to the cosmic microwave
background radiation including all the SDSS galaxies and quasars in the equatorial slice. Its
co-moving radius is 14.0 Gpc. (Since the size of the universe at the epoch of recombination is
smaller that that a present by a factor of 1 + z = 1090, the true radius of this circle is about
12.84 Mpc.) Slightly beyond the cosmic microwave background in co-moving coordinates is
the Big Bang at a co-moving distance of 14.3 Gpc.
(Imagine a point on the cosmic microwave background circle. Draw a radius around that
point that is tangent to the outer circle labeled Big Bang, as shown in the gure, in other words,
a circle that has a radius equal to the difference in radius between the cosmic microwave
background circle and the Big Bang circle. That circle has a co-moving radius of 283 Mpc.
That is the co-moving horizon radius at recombination. If the Big Bang model without
ination were correct we would expect a point on the cosmic microwave background circle
to be causally inuenced only by things inside that horizon radius at recombination. The
angular radius of this small circle as seen from the Earth is (283 Mpc/14, 000 Mpc) radians
or 1.16 degrees. If the Big Bang model without ination were correct we would expect the
cosmic microwave background to be correlated on scales of at most 1.16 degrees. Ination, by
having a short period of accelerated expansion during the rst 10
34
seconds of the universe,
puts distant regions in causal contact because of the slight additional time allowed when the
universe was very small. So, with ination, we can understand why the cosmic microwave
background is uniform to one part in 100,000 all over the sky. Furthermore, random quantum
uctuations predicted by ination add a series of adiabatic uctuations which are expected to
have a peak in the power spectrum at an angular scale about the size of the horizon radius at
recombination calculated above, 0.86 degrees.)
Beyond the Big Bang circle is the circle showing the future co-moving visibility limit. If
we wait until the innite future, we will be able to see out to this circle. (In other words, in the
innite future, we will be able to see particles at the future co-moving visibility limit as they
appeared at the Big Bang.)
The SDSS quasars extend out about halfway out to the cosmic microwave background
radiation. The distribution of quasars shows several features. The radial distribution shows
several shelves due to selection effects as different spectral features used to identify quasars
come into view in the visible. Several radial spokes appear due to incompleteness in some
narrow right ascension intervals. Two large fan shaped regions are empty and not surveyed
because they cover the zone of avoidance close to the galactic plane which is not included
4.4 The Present Universe with Dark Energy 117
Figure 4.8: Galaxies and quasars in the equatorial slice (2 < < 2 degrees) of the Sloan
Digital Sky Survey displayed in co-moving coordinates out to the horizon. The co-moving
distances to galaxies are calculated from measured redshift, assuming Hubble ow and WMAP
cosmological parameters. This is a conformal map it preserves shapes. While this map can
conformally show the complete Sloan survey, the majority of interesting large scale structure is
crammed into a blob in the center. The dashed circle marks the outer limit of gure 4.9. The
circle labeled Unreachable marks the distance beyond which we cannot reach (i.e. we cannot
reach with light signals any object that is further away). This radius corresponds to a redshift
of z = 1.69. As Future comoving visibility limit we label the co-moving distance to which a
photon would travel from the inationary Big Bang to the innite future. This is the maximum
radius out to which observations will ever be possible. At 4.50R
H
0
, it is suprisingly close. [Gott
et al. 2003]
118 4 The Universe with Matter and Dark Energy
in the Sloan survey. These excluded regions run from approximately 3.7 h 8.7 h
and approximately 16.7 h 20.7 h. The quasars do not show noticeable clustering or
large scale structure. This is because the quasars are so widely spaced that the mean distance
between quasars is larger than the correlation length at that epoch.
The circle of reachability is also shown. Quasars beyond this circle are unreachable. Radio
signals emitted by us now will only reach out as far as this circle, even in the innite future.
Figure 4.9: Zoom in of the region marked by the dashed circle in gure 4.8, out to 0.06 r
horizon
(= 858 Mpc). The points shown are galaxies from the main and bright red galaxy samples of the
SDSS. Compared to gure 4.8, we can now see a lot of interesting structure. The Sloan Great
Wall can be seen stretching from 8.7h to 14h in R.A. at a median distance of about 310 Mpc.
Although the large scale structure is easier to see, a zoom in like this fails to capture and
display, in one map, the sizes of modern redshift surveys.
4.4 The Present Universe with Dark Energy 119
The SDSS galaxies appear as a black blob in the center. There is much interesting large
scale structure here but the eld is too crowded and small to show it. This illustrates the
problem of scale in depicting the universe. If we want a map of the entire observable universe
on one page, at a nice scale, the galaxies are crammed into a blob in the center. Let us enlarge
the central circle of radius 0.06 times the distance to the Big Bang circle by a factor of 16.6
and plot it again in gure 4.9. This now shows a circle of co-moving radius 858 Mpc. Almost
all of these points are galaxies from the galaxy and bright red galaxy samples of the SDSS.
Now we can see a lot of interesting structure. The most prominent feature is a Sloan Great
Wall at a median distance of about 310 Mpc stretching from 8.7h to 14h in R.A. There are
numerous voids. A particularly interesting one is close in at a co-moving distance of 125 Mpc
at 1.5h R.A. At the far end of this void are a couple of prominent clusters of galaxies which
are recognizable as ngers of God pointing at the Earth. Redshift in this map is taken as
the co-moving distance indicator assuming participation in the Hubble ow, but galaxies also
have peculiar velocities and in a dense cluster with a high velocity dispersion this causes the
distance errors due to these peculiar velocities to spread the galaxy positions out in the radial
direction producing the nger of God pointing at the Earth. Numerous other clusters can
be similarly identied. This is a conformal map, that preserves shapes excluding the small
effects of peculiar velocities. The original CfA survey in which Geller and Huchra discovered
the Great Wall had a co-moving radius of only 211 Mpc, which is less than a quarter of the
radius shown in gure 4.9. Figure 4.9 is a quite impressive picture, but it does not capture
all of the Sloan Survey. If we displayed gure 4.8 at a scale enlarged by a factor of 16.6 the
central portion of the map would be as you see displayed at the scale shown in gure 2 which
is adequate, but the Big Bang circle would have a diameter of 6.75 feet. You could put this
on your wall, but if we were to print it in the journal for you to assemble it would require the
next 256 pages. This points out the problem of scale for even showing the Sloan Survey all
on one page. Small scales are also not represented well. The distance to the Virgo Cluster in
gure 4.9 is only about 2 mm and the distance from the Milky Way to M31 is only 1/13th of
a millimeter and therefore invisible on this Map. Figure 4.9, dramatic as it is, fails to capture
a picture of all the external galaxies and quasars. The nearby galaxies are too close to see and
the quasars are beyond the limits of the page.
We may try plotting the Universe in lookback time rather than co-moving coordinates.
The result is in gure 4.10. The outer circle is the cosmic microwave background. It is
indistinguishable from the Big Bang as the two are separated by only 380, 000 years out
of 13.7 billion years. The SDSS quasars now extend back nearly to the cosmic microwave
background radiation (since it is true that we are seeing back to within a billion years of the
Big Bang). Lookback time is easier to explain to a lay audience than co-moving coordinates
and it makes the SDSS data look more impressive, but it is a misleading portrayal as far
as shapes and the geometry of space are concerned. It misleads us as to how far out we are
seeing in space. For that, co-moving coordinates are appropriate. gure 4.10 does not preserve
shapes it compresses the large area between the SDSS quasars and the cosmic microwave
background into a thin rim. This is not a conformal map. The SDSS galaxies now occupy
a larger space in the center, but they are still so crowded together that one can not see the
large scale structure clearly. Figure 4.11 shows the central 0.2 radius circle (shown as a dotted
circle in gure 4.10) enlarged by a factor of 5. Thus if we were to make a wall map of the
observable universe using lookback time at the scale of gure 4.11 it would only need to
be 2 feet across and would only require the next 25 pages in the journal to plot. This is an
advantage of the lookback time map. It makes the interesting large scale structure that we see
120 4 The Universe with Matter and Dark Energy
Figure 4.10: Galaxies and quasars in the equatorial slice of the SDSS, displayed in lookback
time coordinates. The radial distance in the gure corresponds to lookback time. While the
Galaxies at the center occupy a larger area, this map is a misleading portrayal as far as shapes
and the geometry of space are concerned. It is not conformal it compresses the area close to
the horizon (this compression is more explicitly shown in gure 4.12). Also, the galaxies are
still too crowded in the center of the map to show all of the intricate details of their clustering.
Figure 4.11 shows a zoom in of the region inside the dashed circle. [Gott et al. 2003]
locally (gure 4.11) a factor of slightly over 3 larger in size relative to the cosmic microwave
background circle than if we had used co-moving coordinates. Figure 4.11 looks quite similar
to gure 4.9. At co-moving radii less than 858 Mpc, the lookback time and co-moving radius
are rather similar. Still, gure 4.11 is not perfectly conformal. Near the outer edges there is
4.4 The Present Universe with Dark Energy 121
a slight radial compression that is beginning to occur in the lookback time map as one goes
toward the Big Bang.
Figure 4.11: Zoom in of the region marked by the dotted circle in gure 4.10, showing SDSS
galaxies out to 0.2 t
horizon
. The details of galaxy clustering are now displayed much better.
However, like gure 4.9, it still fails to capture the whole survey in one, reasonably sized, map.
[Gott et al. 2003]
The effects of radial compression are illustrated in gure 4.12, where we have plotted
a square grid in co-moving coordinates in terms of lookback time as would be depicted in
gure 4.10. Each grid square would contain an equal number of galaxies in a at slice of
constant vertical thickness. This shows the distortion of space that is produced by using the
lookback time. The squares become more and more distorted in shape as one approaches the
edge.
122 4 The Universe with Matter and Dark Energy
Figure 4.12: Square comoving grid shown in lookback time coordinates. Grid spacing is
0.1R
H
0
= 422.24 Mpc. Each grid square would contain an equal number of galaxies in a
at slice of constant vertical thickness. The distortion of space that is produced by using the
lookback time is obvious as the squares become more and more distorted in shape as one ap-
proaches the horizon. [Gott et al. 2003]
4.5 Measuring the Acceleration of the Present Universe
The three primary methods to measure curvature are luminosity, scale length and number.
Luminosity requires an observer to nd some standard candle, such as the brightest quasars or
Supernovae, and follow them out to high redshifts. Scale length requires that some standard
size be used, such as the size of the largest galaxies. Lastly, number counts are used where
one counts the number of galaxies in a box as a function of distance. Several groups are
measuring distant supernovae with the goal of determining whether the Universe is open or
closed by measuring the curvature in the Hubble diagram. The gure 4.14 shows a binned
version of the latest dataset (Kowalski et al. (2008)).
4.5 Measuring the Acceleration of the Present Universe 123
4.5.1 Luminosity Distance and HubbleDiagrams for CDM
With standard candles, one can measure the cosmos over the distance modulus
m(z) M = 5 log d
L
(z;
M
,

, H
0
) + 25 , (4.187)
which connects apparent magnitude m(z) und absolute magnitude M with the distance d
L
in units of Mpc. Hereby, d
L
is the luminosity distance d
L
= r
1
R
0
(1 + z), where r
1
is the
coordinate distance of the object at redshift z
R
0
_
r
1
0
dr

1 kr
2
= R
0
_
t
0
t
1
c dt
R(t)
=
c
H
0
_
t
0
t
1
da
a
_

M
/a +

a
2
+ 1
M

. (4.188)
The second equality follows from the Friedman equation. Since da = dz/(1 + z)
2
, we
d_L [Gpc]
1 10
z
10
-1
1 Vacuum
Luminosity Distance Inflationary Universe
Figure 4.13: Luminosity distance as a function of redshift in the at inationary Universe. The
linear relation corresponds to the classical Hubble law (or a closed Universe with
M
= 2,

= 0), the boxes to a matterdominated model


M
= 1 (SCDM), the dotted curve to an
open model
M
= 0.3,

= 0, the star symbols refer to a CDMModell


M
= 0.3,

= 0.7 (CDM), the lowest curve to a vacuum dominated de Sitter Universe with

= 1
(H
0
= 70 km/s/Mpc).
have the equation
R
0
dr

1 kr
2
=
c
H
0
dz
E(z)
(4.189)
124 4 The Universe with Matter and Dark Energy
with
E(z)
_

k
(1 +z)
2
+
M
(1 +z)
3
+

(4.190)
and E(0) = 1. This denes the comoving distance d
C
(z) = r
1
R
0
, where r
1
has to be the
solution of
R
0
_
r
1
0
dr

1 kr
2
=
1

0
R
0
arcsin(h)(

0
r
1
) =
c
H
0
_
z
0
dz

E(z

)
. (4.191)
For our analysis we use the luminosity distance d
L
(z) = (1 +z)d
C
(z) in the general form of
d
L
(z) =
c
H
0
1 +z

0
S
_

0
_
z
0
dz

_
i

i
(1 +z

)
3+3w
i

0
(1 +z

)
2
_
. (4.192)
S(x) = sin(x), x, or sinh(x) for closed, at, and open models respectively, and the curvature
parameter
0
, is dened as
0
=
k
=

i

i
1. This formula includes the special case
of a constant vacuum contribution with w
V
= 1.
The Hubble parameter H(z), appearing in these equations is provided by the Einstein eld
equations. If the different sources which populate the universe do not interact with each other
and each of them is represented by an equation of state w
i
P
i
/
i
(which can be a function
of time in general), energy conservation
d
dz
= 3
1 +w(z)
1 +z
(4.193)
leads to the solution for the density as a function of redshift
(z) =
0
exp
_
3
_
z
0
1 +w
i
(z

)
1 +z

dz

_
. (4.194)
The Friedman equations then yield the solution for the Hubble parameter
H
2
(z) = H
2
0
_

i
exp
_
3
_
z
0
1 +w
i
(z

)
1 +z

dz

k
(1 +z)
2
_
, (4.195)
and for the curvature
q(z) =
H
2
0
H
2
(z)

i
2
{1 + 3w
i
(z)} exp
_
3
_
z
0
1 +w
i
(z

)
1 +z

dz

_
, (4.196)
where
i
are, as usual, the present day energy densities of the different source components in
units of the critical density 3H
2
0
/8G and
k
k/R
2
0
H
2
0
(i denoting non-relativistic matter,
radiation, cosmological constant, quintessence etc.). The present value of the scale factor R
0
,
which measures the curvature of spacetime, can now be calculated from
R
0
= H
1
0

k
(

i

i
1)
. (4.197)
4.5 Measuring the Acceleration of the Present Universe 125
Figure 4.14: The curves show a closed Universe ( = 2, classical Hubble law) in red, the
critical density Universe ( = 1, SCDM) in black, the empty Universe ( = 0) in green, the
steady state model in blue, and the WMAP based concordance model with
m
= 0.27 and

DE
= 0.73 in purple. This model gives H
0
= 71 km/sec/Mpc which has been used to scale
the luminosity distances in the plot. The data show an accelerating Universe at low to moderate
redshifts but a decelerating Universe at higher redshifts, consistent with a model having both a
cosmological constant and a signicant amount of dark matter. The dashed black curve shows
an Einstein-de Sitter model with a constant co-moving dust density which can be ruled out. The
dashed purple curve shows a closed LCDM model which is a good t to the data. The dashed
blue curve shows an evolving supernova model which is also a good t. [Gold Sample: Ned
Wright 2008]
We note that the coordinate distance r
1
, and hence d
L
, are sensitive to
i
for the distant SNe
only. For the nearby SNe (in the low-redshift limit), equation (4.187) then reduces to
m(z) = M+ 5 log z , (4.198)
which can be used to measure Mby using low-redshift supernovae-measurements that are far
enough into the Hubble ow so that their peculiar velocities do not contribute signicantly to
their redshifts.
126 4 The Universe with Matter and Dark Energy
Special Cases
For dark matter (i = M, w
M
= 0) and a vacuum energy (i = V , w
V
= 1), the integral can
be reduced to the form
d
L
(z;
M
,

, H
0
) =
c(1 +z)
H
0
_
|
k
|
S(x[z]) , (4.199)
where
x[z] =
_
|
k
|
_
z
0
dz

_
(1 +z

)
2
(1 +
M
z

) z

(2 +z

. (4.200)
The curvature parameter
k
follows from the condition (4.95). This integral is easily com-
puted numerically. From this we get the apparent magnitude for given absolute magnitude M
and therefore a reduced distance modulus which does not depend on the Hubble constant
DM m(z) M= 5 log D
L
(z;
M
,

) (4.201)
with d
L
(c/H
0
) D
L
and the denition of
M M 5 log
_
H
0
Mpc
c
_
+ 25 = M 5 log h + 42.38 , (4.202)
which is constant within an ensemble of standard candles. From tting lowredshift SNe,
m
B
(z) = M
B
+5 log z, one can derive a value for M. Perlmutter et al. have pointed out in
1997 that the apparent magnitude of SNIa at redshifts z > 0.1 only depends on the function
D
L
(z;
M
,

), which does not contain the Hubble constant H


0
(Fig. 4.16). This idea is the
basic driver behind all high redshift supernovae projects.
For a at Universe,
k
= 0, the above relation is simplied to
d
L
(z;
M
,

, H
0
) =
c(1 +z)
H
0
_
z
0
dz

M
(1 +z

)
3
+ 1
M
. (4.203)
The special case with
M
= 1, i.e.

= 0, reduces to the well known luminosity distance


for at Universe without vacuum energy
d
L
(z; H
0
) =
2c
H
0
_
1 +z

1 +z

. (4.204)
In general, the objects appear dimmer in a Universe including vacuum energy when compared
to classical CDM models. Please note that this is also the case for Quasars. At high redshifts,
this difference can amount upto two magnitudes (Fig. 4.15).
The KCorrection
When we measure magnitudes of cosmological objects, we have to take into account that the
spectrum is shifted with respect to the measured wavelengths due to cosmic expansion. Let us
consider a quasar with redshift z, which has emitted its light at time t
1
. Let E

(t) denote the


4.5 Measuring the Acceleration of the Present Universe 127
Redshift
0 1 2 3 4 5
b
r
i
g
h
t
























D
M

















f
a
i
n
t
-8
-6
-4
-2
0
2
4
6
8
Vacuum
Closed Universe
Reduced Distance Modulus
Figure 4.15: Reduced distance modulus DM = m M as function of redshift for various
cosmological models. For high redshift Supernovae we see differences upto two magnitudes.
The starry symbols represent a model with
M
= 0.3 and

= 0.7. The vacuum model has

= 1, the closed Universe corresponds to


M
= 2 and

= 0 (global linear Hubblelaw).


The dotted curve is an open model with
M
= 0.3 and

= 0, and the boxes relate to the


classical SCDM. In this redshiftrange, an open model is not very much different from LCDM.
monochromatic luminosity for wavelength and time t, measured in the rest system of the
quasar. Let L

0
be the measured luminosity in a spectral band centered at
0
(in erg/s/cm
2
)
L

0
=
_

0
F

d. (4.205)
S

denotes the transmission function of the instrument. We nd then


3
L

0
=
1
4d
2
(1 +z)
_

0
E
/(1+z)
(t
1
) S

d
=
_

0
E

(t
0
) S

d
4d
2
(1 +z)

_

0
E
/(1+z)
(t
0
) S

d
_

0
E

(t
0
) S

0
E
/(1+z)
(t
1
) S

d
_

0
E
/(1+z)
(t
0
) S

d
. (4.206)
3
Tinsley, B.M. 1970, Ap&SS 6, 344
128 4 The Universe with Matter and Dark Energy
Figure 4.16: Hubble Diagram for Supernovae Type Ia. Top panel: Hubble Diagram for SCP
lowextinction subsample; Bottom panel: Residuals relative to an empty Universe. Data from
Knop et al. 2003 [9].
with d as the luminosity distance of the quasar. From this expression we obtain the apparent
magnitude
m

0
= M

0
(t
0
) + 5 log d +const
+
_
2.5 log(1 +z) + 2.5 log
_

0
E

(t
0
) S

d
_

0
E
/(1+z)
(t
0
) S

d
_
+ 2.5 log
_
_

0
E
/(1+z)
(t
0
) S

d
_

0
E
/(1+z)
(t
1
) S

d
_
(4.207)
4.5 Measuring the Acceleration of the Present Universe 129
The expression in the brackets denotes the Kcorrection, the last term corresponds to evolu-
tionary effects. The observed magnitude consists therefore of 5 different contributions D
the absolut luminosity in the rest system of the object;
a factor depending on the luminosity distance;
a constant factor from normalisation;
the Kcorrection the difference in magnitudes between two objects of the same spec-
trum, however shifted with respect to each other (the difference between observed mag-
nitude at wavelength
1
=
0
/(1 + z) and the magnitude of the same quasar at time t
0
with wavelength
0
);
evolutionary effects (ageing of quasars and galaxies).
4.5.2 Measuring Cosmology with Supernovae
Understanding the global history of the Universe is a fundamental goal of cosmology. One
of the conceptually simplest tests in the repertoire of the cosmologist is observing how a
standard candle dims as a function of redshift. The nearby Universe provides the current rate
of expansion, and with more distant objects it is possible to start seeing the varied effects
of cosmic curvature and the Universes expansion history (usually expressed as the rate of
acceleration/deceleration). Over the past several decades a paradigm for understanding the
global properties of the Universe has emerged based on General Relativity with the assumption
of a homogeneous and isotropic Universe. The relevant constants in this model are the Hubble
constant (or current rate of cosmic expansion), the relative fractions of species of matter that
contribute to the energy density of the Universe, and these species equation of state.
Early luminosity distance investigations used the brightest objects available for measuring
distance - bright galaxies (Fig. 4.17), but these efforts were hampered by the impreciseness
of the distance indicators and the changing properties of the distance indicators as a function
of lookback time. Although many other methods for measuring the global curvature and
cosmic deceleration exist, supernovae (SNe) have emerged as one of the preeminent distance
methods due to their signicant intrinsic brightness (which allows them to be observable in
the distant Universe), ubiquity (they are visible in both the nearby and distant Universe), and
their precision (type Ia SNe provide distances that have a precision of approximately 8%).
Over the past decade, supernovae have emerged as some of the most powerful tools for
measuring extragalactic distances. A well developed physical understanding of type II super-
novae allow them to be used to measure distances independent of the extragalactic distance
scale. Type Ia supernovae are empirical tools whose precision and intrinsic brightness make
them sensitive probes of the cosmological expansion. Both types of supernovae are consistent
with a Hubble Constant within 10% of H
0
= 70 km/s/Mpc. Two teams have used type Ia
supernovae to trace the expansion of the Universe to a look-back time more than 60% of the
age of the Universe. These observations show an accelerating Universe which is currently best
explained by a cosmological constant or other form of dark energy with an equation of state
near w = P/ = 1. While there are many possible remaining systematic effects, none ap-
pears large enough to challenge these current results. Future experiments are planned to better
characterize the equation of state of the dark energy leading to the observed acceleration by
observing hundreds or even thousands of objects. These experiments will need to carefully
130 4 The Universe with Matter and Dark Energy
Figure 4.17: Hubble diagram for the brightest galaxies in clusters (elliptical galaxies). Magni-
tudes are corrected for various effects. [Source: Sandage].
control systematic errors to ensure future conclusions are not dominated by effects unrelated
to cosmology.
Type Ia Supernovae as Standardized Candles
SNIa have been used as extragalactic distance indicators since Kowal rst published his Hub-
ble diagram ( = 0.6 mag) for type I SNe. We now recognize that the old type I SNe spec-
troscopic class is comprised of two distinct physical entities: SNIb/c which are massive stars
that undergo core collapse (or in some rare cases might undergo a thermonuclear detonation
in their cores) after losing their hydrogen atmospheres, and SNIa which are most likely ther-
monuclear explosions of white dwarfs. In the mid-1980s it was recognized that studies of the
type I SN sample had been confused by these similar appearing SNe, which were henceforth
classied as type Ib and type Ic. By the late 1980s/early 1990s, a strong case was being made
that the vast majority of the true type Ia SNe had strikingly similar light curve shapes, spectral
time series, and absolute magnitudes. A 1992 review by Branch and Tammann [2] of a variety
of studies in the literature concluded that the intrinsic dispersion in B and V maximum for
type Ia SNe must be < 0.25 mag, making them the best standard candles known so far.
In fact, the Branch and Tammann review indicated that the magnitude dispersion was
probably even smaller, but the measurement uncertainties in the available datasets were too
4.5 Measuring the Acceleration of the Present Universe 131
large to tell. The Calan/Tololo Supernova Search (CTSS), a program begun by Hamuy et al.
in 1990, took the eld a dramatic step forward by obtaining a crucial set of high quality SN
light curves and spectra. By targeting a magnitude range that would discover type Ia SNe in
the redshift range z = 0.01 0.1, the CTSS was able to compare the peak magnitudes of
SNe whose relative distance could be deduced from their Hubble velocities. As the CTSS
data began to become available, several methods were presented that could select for the most
standard subset of the type Ia standard candles, a subset which remained the dominant majority
of the ever-growing sample. Phillips [?] found a tight correlation between the rate at which
a type Ia SNs luminosity declines and its absolute magnitude, a relation which apparently
applied not only to the Branch Normal type Ia SNe, but also to the peculiar type Ia SNe.
Phillips plotted the absolute magnitude of the existing set of nearby SNIa, which had dense
photoelectric or CCD coverage, versus the parameter m
15
(B), the amount the SN decreased
in brightness in the B-band over the 15 days following maximum light. The sample showed
a strong correlation which, if removed, dramatically improved the predictive power of SNIa.
Hamuy et al. [7] used this empirical relation to reduce the scatter in the Hubble diagram to
< 0.2 mag in V for a sample of nearly 30 SNIa from the CTSS search.
Impressed by the success of the m
15
(B) parameter, Riess et al. [17] developed the multi-
color light curve shape method (MLCS), which parameterized the shape of SN light curves as
a function of their absolute magnitude at maximum. This method also included a sophisticated
error model and tted observations in all colors simultaneously, allowing a color excess to be
included. This color excess, which we attribute to intervening dust, enabled the extinction to
be measured. Another method that has been used widely in cosmological measurements with
SNIa is the stretch method described in Perlmutter et al. [14]. This method is based on
the observation that the entire range of SNIa light curves, at least in the B and V bands, can
be represented with a simple time stretching (or shrinking) of a canonical light curve. The
coupled stretched B and V light curves serve as a parameterized set of light curve shapes,
providing many of the benets of the MLCS method but as a much simpler (and constrained)
set. This method, as well as recent implementations of m
15
(B), also allows extinction to be
directly incorporated into the SNIa distance measurements. Other methods that correct for
intrinsic luminosity differences or limit the input sample by various criteria have also been
proposed to increase the precision of type Ia SNe as distance indicators. While these latter
techniques are not as developed as the m
15
(B), MLCS, and stretch methods, they all provide
distances that are comparable in precision, roughly = 0.18 mag about the inverse square law,
equating to a fundamental precision of SNIa distances of 6% (0.12 mag), once photometric
uncertainties and peculiar velocities are removed. Finally, a poor mans distance indicator, the
snapshot method, combines information contained in one or more SN spectra with as little as
one nights multi-color photometry. This methods accuracy depends critically on how much
information is available.
To illustrate the effect of cosmological parameters on the luminosity distance, in Fig. 4.19
we plot a series of models. In the top panel, the various models show the same linear behavior
at z < 0.1 with models having the same H
0
being indistinguishable to a few percent. By
z = 0.5 the models with signicant are clearly separated, with luminosity distances that are
signicantly further than the zero- universes. Unfortunately, two perfectly reasonable uni-
verses, given our knowledge of the local matter density of the Universe (M 0.2), one with
a large cosmological constant,

= 0.7,
M
= 0.3 and one with no cosmological constant,

M
= 0.2, show differences of less than 25%, even to redshifts of z > 5. Interestingly, the
maximum difference between the two models is at z 0.8, not at large z. Fig. 4.20 illus-
132 4 The Universe with Matter and Dark Energy
Figure 4.18: The relationship between the light curve width and the luminosity of SNIa. The
top panel shows various light curves from the Calan/Tololo SN survey. The bottom graph shows
how stretch can be used to describe the light curve with just one parameter.
4.5 Measuring the Acceleration of the Present Universe 133
Figure 4.19: d
L
expressed as distance modulus (mM) for four relevant cosmological models;

M
= 0,

= 0 (empty Universe, solid line);


M
= 0.3,

= 0 (short dashed line);

M
= 0.3,

= 0.7 (hatched line); and


M
= 1.0,

= 0 (long dashed line). In the bottom


panel the empty Universe has been subtracted from the other models to highlight the differences.
trates the effect of changing the equation of state of the non-matter, dark energy component,
assuming a at universe, tot = 1. If we are to discern a dark energy component that is not a
cosmological constant, measurements better than 5% are clearly required, especially since the
differences in this diagram include the assumption of atness and also x the value of
M
. In
fact, to discriminate among the full range of dark energy models with time varying equations
of state will require much better accuracy than even this challenging goal.
The intrinsic brightness of SNIa allow them to be discovered to z > 1.5 with current
instrumentation (while a comparably deep search for type II SNe would only reach redshifts
of z 0.5). In the 1980s, however, nding, identifying, and studying even the impressively
luminous type Ia SNe was a daunting challenge, even towards the lower end of the redshift
range shown in Fig. 4.19. At these redshifts, beyond z 0.25, Fig. 4.19 shows that relevant
cosmological models could be distinguished by differences of order 0.2 mag in their predicted
luminosity distances. For SNIa with a dispersion of 0.2 mag, 10 well observed objects should
provide a 3 separation between the various cosmological models. It should be noted that the
uncertainty described above in measuring H
0
is not important in measuring the parameters for
different cosmological models. Only the relative brightness of objects near and far is being
exploited and the absolute value of H
0
scales out.
134 4 The Universe with Matter and Dark Energy
Figure 4.20: d
L
for a variety of cosmological models containing
M
= 0.3 and
x
= 0.7 with
a constant (not time-varying) equation of state w
x
. The w
x
= 1 model has been subtracted
off to highlight the differences between the various models.
The rst distant SN search was started by the Danish team of NorgaardNielsen et al.
With signicant effort and large amounts of telescope time spread over more than two years,
they discovered a single SNIa in a z = 0.3 cluster of galaxies (and one SNII at z = 0.2).
The SNIa was discovered well after maximum light on an observing night that could not have
been predicted, and was only marginally useful for cosmology. However, it showed that such
high redshift SNe did exist and could be found, but that they would be very difcult to use as
cosmological tools.
Just before this rst discovery in 1988, a search for high redshift type Ia SNe using a
then novel wide eld camera on a much larger (4m) telescope was begun at the Lawrence
Berkeley National Laboratory (LBNL) and the Center for Particle Astrophysics, at Berkeley.
This search, now known as the Supernova Cosmological Project (SCP), was inspired by the
impressive studies of the late 1980s indicating that extremely similar type Ia SN events could
be recognized by their spectra and light curves, and by the success of the LBNL fully robotic
4.5 Measuring the Acceleration of the Present Universe 135
low-redshift SN search in nding 20 SNe with automatic image analysis.
The SCP targeted a much higher redshift range, z > 0.3, in order to measure the (pre-
sumed) deceleration of the Universe, so it faced a different challenge than the CTSS search.
The high redshift SNe required discovery, spectroscopic conrmation, and photometric follow
up on much larger telescopes. This precious telescope time could neither be borrowed from
other visiting observers and staff nor applied for in sufcient quantities spread throughout the
year to cover all SNe discovered in a given search eld, and with observations early enough
to establish their peak brightness. Moreover, since the observing time to conrm high redshift
SNe was signicant on the largest telescopes, there was a clear chicken and egg problem:
telescope time assignment committees would not award follow-up time for a SN discovery
that might, or might not, happen on a given run (and might, or might not, be well past max-
imum) and, without the follow-up time, it was impossible to demonstrate that high redshift
SNe were being discovered by the SCP.
By 1994, the SCP had solved this problem, rst by providing convincing evidence that
SNe, such as SN1992bi, could be discovered near maximum (and K-corrected) out to z =
0.45, and then by developing and successfully demonstrating a new observing strategy that
could effectively guarantee SN discoveries on a predetermined date, all before or near maxi-
mum light. Instead of discovering a single SN at a time on average (with some runs not nding
one at all), the new approach aimed to discover an entire batch of half-a-dozen or more type
Ia SNe at a time by observing a much larger number of galaxies in a single two or three day
period a few nights before new Moon. By comparing these observations with the same obser-
vations taken towards the end of dark time almost three weeks earlier, it was possible to select
just those SNe that were still on the rise or near maximum. The chicken and egg problem was
solved, and now the follow-up spectroscopy and photometry could be applied for and sched-
uled on a pre-specied set of nights. The new strategy worked - the SCP discovered batches
of high redshift SNe, and no one would ever again have to hunt for high-redshift SNe without
the crucial follow-up scheduled in advance.
The High-Z SN Search (HZSNS) was conceived at the end of 1994, when this group of
astronomers became convinced that it was both possible to discover SNIa in large numbers at
z > 0.3 by the efforts of Perlmutter et al., and also use them as precision distance indicators as
demonstrated by the CTSS group. Since 1995, the SCP and HZSNS have both worked avidly
to obtain a signicant set of high redshift SNIa.
The two high redshift teams both used this pre-scheduled discovery and follow-up batch
strategy. They each aimed to use the observing resources they had available to best scientic
advantage, choosing, for example, somewhat different exposure times or lters. Quantita-
tively, type Ia SNe are rare events on an astronomers time scale - they occur in a galaxy like
the Milky Way a few times per millennium and the chapter by Cappellaro in this volume).
With modern instruments on 4 meter-class telescopes, which observe 1/3 of a square degree
to R = 24 mag in less than 10 minutes, it is possible to search a million galaxies to z < 0.5
for SNIa in a single night.
Since SNIa take approximately 20 days to rise from undetectable to maximum light, the
three-week separation between observing periods (which equates to 14 rest frame days at
z = 0.5) is a good lter to catch the SNe on the rise. The SNe are not always easily identied
as new stars on the bright background of their host galaxies, so a relatively sophisticated
process must be used to identify them. The process, which involves 20 Gigabytes of imaging
data per night, consists of aligning a previous epoch, matching the image star proles (through
convolution), and scaling the two epochs to make the two images as identical as possible. The
136 4 The Universe with Matter and Dark Energy
difference between these two images is then searched for new objects which stand out against
the static sources that have been largely removed in the differencing process. The dramatic
increase in computing power in the 1980s was an important element in the development of
this search technique, as was the construction of wide-eld cameras with ever larger CCD
detectors or mosaics of such detectors.
This technique is very efcient at producing large numbers of objects that are, on average,
at or near maximum light, and does not require unrealistic amounts of large telescope time.
It does, however, place the burden of work on follow-up observations, usually with different
instruments on different telescopes. With the large number of objects discovered (50 in two
nights being typical), a new strategy is being adopted by both the SCP and HZSNS teams,
as well as additional teams like the Canada France Hawaii Telescope (CFHT) legacy survey,
where the same elds are repeatedly scanned several times per month, in multiple colors, for
several consecutive months. This type of observing program provides both discovery of new
objects and their follow up, all integrated into one efcient program. It does require a large
block of time on a single telescope - a requirement which was not politically feasible in years
past, but is now possible.
Obstacles to Measuring Luminosity Distances at High-Z
As shown above, the distances measured to SNIa are well characterized at z < 0.1, but com-
paring these objects to their more distant counterparts requires great care. Selection effects
can introduce systematic errors as a function of redshift, as can uncertain K-corrections and a
possible evolution of the SNIa progenitor population as a function of look-back time. These
effects, if they are large and not constrained or corrected, will limit our ability to accurately
measure relative luminosity distances, and have the potential to reduce the efcacy of high-z
type Ia SNe for measuring cosmology
Kcorrection: As SNe are observed at larger and larger redshifts, their light is shifted
to longer wavelengths. Since astronomical observations are normally made in xed band
passes on Earth, corrections need to be applied to account for the differences caused
by the spectrum shifting within these band passes. These corrections take the form of
integrating the spectrum of an SN over the relevant band passes, shifting the SN spectrum
to the correct redshift, and re-integrating.
Extinction: In the nearby Universe we see SNIa in a variety of environments, and about
10% have signicant extinction. Since we can correct for extinction by observing two or
more wavelengths, it is possible to remove any rst order effects caused by a changing
average extinction of SNIa as a function of z. However, second order effects, such as
possible evolution of the average properties of intervening dust, could still introduce
systematic errors. This problem can also be addressed by observing distant SNIa over a
decade or so of wavelength in order to measure the extinction law to individual objects.
Unfortunately, this is observationally very expensive. Current observations limit the total
systematic effect to < 0.06 mag, as most of our current data is based on two color
observations.
An additional problem is the existence of a thin veil of dust around the Milky Way.
Measurements from the Cosmic Background Explorer (COBE) satellite accurately deter-
mined the relative amount of dust around the Galaxy, but there is an uncertainty in the
4.5 Measuring the Acceleration of the Present Universe 137
absolute amount of extinction of about 2 - 3%. This uncertainty is not normally a prob-
lem, since it affects everything in the sky more or less equally. However, as we observe
SNe at higher and higher redshifts, the light from the objects is shifted to the red and
is less affected by the Galactic dust. Our present knowledge indicates that a systematic
error as large as 0.06 mag is attributable to this uncertainty.
Selection Effects: As we discover SNe, we are subject to a variety of selection effects,
both in our nearby and distant searches. The most signicant effect is the Malmquist Bias
- a selection effect which leads magnitude limited searches to nd brighter than average
objects near their distance limit since brighter objects can be seen in a larger volume than
their fainter counterparts. Malmquist Bias errors are proportional to the square of the
intrinsic dispersion of the distance method, and because SNIa are such accurate distance
indicators these errors are quite small, 0.04 mag. Monte Carlo simulations can be
used to estimate such selection effects, and to remove them from our data sets. The total
uncertainty from selection effects is 0.01 mag and, interestingly, may be worse for
lower redshift objects because they are, at present, more poorly quantied.
Gravitational Lensing: Several authors have pointed out that the radiation from any
object, as it traverses the large scale structure between where it was emitted and where
it is detected, will be weakly lensed as it encounters uctuations in the gravitational
potential. On average, most of the light travel paths go through under-dense regions and
objects appear de-magnied. Occasionally, the light path encounters dense regions and
the object becomes magnied. The distribution of observed uxes for sources is skewed
by this process such that the vast majority of objects appear slightly fainter than the
canonical luminosity distance, with the few highly magnied events making the mean of
all light paths unbiased. Unfortunately, since we do not observe enough objects to capture
the entire distribution, unless we know and include the skewed shape of the lensing a bias
will occur. At z = 0.5, this lensing is not a signicant problem: If the Universe is at in
normal matter, the large scale structure can induce a shift of the mode of the distribution
by only a few percent. However, the effect scales roughly as z2, and by z = 1.5 the effect
can be as large as 25%. While corrections can be derived by measuring the distortion of
background galaxies near the line of sight to each SN, at z > 1, this problem may be
one which ultimately limits the accuracy of luminosity distance measurements, unless
a large enough sample of SNe at each redshift can be used to characterize the lensing
distribution and average out the effect. For the z 0.5 sample, the error is < 0.02 mag,
but it is much more signicant at z > 1, especially if the sample size is small.
Evolution: SNIa are seen to evolve in the nearby Universe. Hamuy et al. [29] plotted the
shape of the SN light curves against the type of host galaxy. SNe in early hosts (galaxies
without recent star formation) consistently show light curves which rise and fade more
quickly than SNe in late-type hosts (galaxies with on-going star formation). However,
once corrected for light curve shape the luminosity shows no bias as a function of host
type. This empirical investigation provides reassurance for using SNIa as distance indica-
tors over a variety of stellar population ages. It is possible, of course, to devise scenarios
where some of the more distant SNe do not have nearby analogues, so as supernovae
are studied at increasingly higher redshifts it can become important to obtain detailed
spectroscopic and photometric observations of every distant SN to recognize and reject
examples that have no nearby analogues.
138 4 The Universe with Matter and Dark Energy
Figure 4.21: Upper panel: The Hubble diagram for high redshift SNIa from both the HZSNS
and the SCP teams. Lower panel: The residual of the distances relative to a
M
= 0.3,

=
0.7 Universe. The z < 0.15 objects for both teams are drawn from CTSS sample, so many
of these objects are in common between the analyses of the two teams. Lowest panel: In the
last few years distant supernovae with redshifts up to 1.755 have been observed by the Hubble
Space Telescope. These objects show that the trend toward fainter supernovae seen at moderate
redshifts has reversed. The green curve is the = 0 Universe. The solid magenta curve shows
the best t at accelerating vacuum-dominated model. The dashed magenta curve is the best
closed dark energy dominated t to the supernova data alone.
4.5 Measuring the Acceleration of the Present Universe 139
High Redshift SNIa Observations
The SCP [?] in 1997 presented their rst results with 7 objects at a redshift around z = 0.4.
Figure 4.22: The supernova data as of April 2008 published by Kowalski et al. (2008) provide
the best t, 1, 2 and 3 standard deviation contours shown as the green, blue, red and black ellipses
in the gure at left. The CMB data using WMAP ve year results provide the cloud of dots from
a Monte Carlo Markov chain sampling of the likelihood function. The CMB degeneracy track
does not follow the at Universe line, but crosses the at line at a point reasonably consistent
with the supernova t. Each CMB model has an implied Hubble constant which provides the
color code for the dots. A model that ts both the supernova data and the CMBdata has a Hubble
constant that agrees reasonably well with the Hubble Space Telescope Key Project value of the
Hubble constant. The addition of high redshift supernovae has had two effects on the supernova
error ellipse. The long axis of the ellipse has gotten shorter, and the slope of the ellipse has
gotten higher. The best t model has gotten closer to the CMB degeneracy track in absolute
terms, and it has also gotten closer in terms of standard deviations in the Kowalski et al. (2008)
dataset. [Plot: Ned Wright 2008]
These objects hinted at a decelerating Universe with a measurement of
M
= 0.88
+0.69
0.60
,
but were not denitive. Soon after, the SCP published a further result, with a z 0.84 SNIa
observed with the KECK I and HST added to the sample, and the HZSNS presented the results
from their rst four objects. The results from both teams now ruled out a
M
= 1 Universe
with greater than 95% signicance. These ndings were again superceded dramatically when
both teams announced results including more SNe (10 more HZSNS SNe, and 34 more SCP
140 4 The Universe with Matter and Dark Energy
SNe) that showed not only were the SN observations incompatible with a
M
= 1 Universe,
they were also incompatible with a Universe containing only normal matter. Fig. 4.21 shows
the Hubble diagram for both teams. Both samples show that SNe are, on average, fainter than
would be expected, even for an empty Universe, indicating that the Universe is accelerating.
The agreement between the experimental results of the two teams is spectacular, especially
considering the two programs have worked in almost complete isolation from each other.
The easiest solution to explain the observed acceleration is to include an additional com-
ponent of matter with an equation of state parameter more negative than w < 1/3; the most
familiar being the cosmological constant (w = 1). Fig. 4.23 shows the joint condence con-
tours for values of
M
and

fromboth experiments. If we assume the Universe is composed


only of normal matter and a cosmological constant, then with greater than 99.9% condence
the Universe has a non-zero cosmological constant or some other form of dark energy. Of
Figure 4.23: Left panel: Contours of
M
versus w
x
from current observational data. Right
Panel: Contours of
M
versus w
x
from current observational data, where the current value of

M
is obtained from the 2dF redshift survey. For both panels
M
+
x
= 1 is taken as a prior.
course, we do not know the form of dark energy which is leading to the acceleration, and it
is worthwhile investigating what other forms of energy are possible additional components.
Fig. 4.23 shows the joint condence contours for the HZSNS+SCP observations for
M
and
w
x
(the equation of state of the unknown component causing the acceleration). Because this
introduces an extra parameter, we apply the additional constraint that
M
+
x
= 1, as in-
dicated by the CMB experiments. The cosmological constant is preferred, but anything with
a w < 0.5 is acceptable. Additionally, we can add information about the value of
M
, as
supplied by recent 2dF redshift survey results, as shown in the 2nd panel, where the constraint
strengthens to w < 0.6 at 95% condence [16].
4.5 Measuring the Acceleration of the Present Universe 141
The mz Fit Procedure
Nowfor given Mand
i
, these equations can provide the predicted value of m(z) at any given
z. We compare this value with the corresponding observed magnitude m
obs
and compute
2
from

2
=
N

j=1
_
m(z
j
; M,
i
) m
obs,j

m
obs,j
_
2
, (4.208)
where the quantity
m
obs,j
is the uncertainty in the observed magnitude m
obs,j
of the j-th SN.
It may be noted that some data sets are given in terms of the distance modulus = m(z)M,
instead of m. However, the zero-point absolute magnitude is set arbitrarily. In this case also
we can use equation (4.208) for tting the data by using
obs
in place of m
obs
, the constant M
(which plays the role of the normalization constant in this case) simply gets modied suitably.
Sometimes in the data we have also been provided with independent uncertainties on some
other variable. In this case the equation for
2
gets modied as

2
=
N

j=1
_
{m(z
j
; M,
i
) m
obs,j
}
2

2
m
obs,j
+
2
int
+
2
v,j
_
, (4.209)
where
int
is the uncertainty due to the intrinsic dispersion of SNe absolute magnitude, and

v,j
due to peculiar velocity.
The key point about the SNe Ia data sets is that the absolute luminosities M of all the SNe,
distant or nearby, are regarded same (standard candle-hypothesis). Hence so is the constant
M, as it has only one extra parameter H
0
which certainly does not differ from SN to SN.
Thus there are two ways of the actual data tting: (i) estimate Mby using low-redshift SNe,
and use this value in equation (4.208) to estimate
i
from the high-redshift data alone; (ii) use
low-, as well as, high-redshift data simultaneously to evaluate all the parameters from equation
(4.208) by keeping Mas a free parameter. Obviously the second method gives a better tting,
as shown in Table 4.2.
It is obvious from equation (4.208) that if the model represents the data correctly, then the
difference between the predicted magnitude and the observed one at each data point should be
roughly the same size as the measurement uncertainties and each data point will contribute to

2
roughly one, giving the sum roughly equal to the number of data points N (more correctly
Nnumber of tted parameters number of degrees of freedom dof). If
2
is large, then
the t is bad. However we must quantify our judgment and decision about the goodness-
of-t, in the absence of which, the estimated parameters of the model (and their estimated
uncertainties) have no meaning at all. An independent assessment of the goodness-of-t of
the data to the model is given in terms of the
2
-probability: if the tted model provides a
typical value of
2
as x at n dof, this probability is given by
P(x, n) =
1
(n/2)
_

x/2
e
u
u
n/21
du. (4.210)
P(x, n) gives the probability that a model which does t the data at n dof, would give a value
of
2
as large or larger than x. If P is very small, the model is ruled out. For example, if we
get a
2
= 20 at 5 dof for some model, then the hypothesis that the model describes the data
genuinely is unlikely, as the probability P(20, 5) = 0.0012 is very small.
142 4 The Universe with Matter and Dark Energy
Fitting with Various SNe Data
In a recent paper, Vishwakarma [24] has discussed the most recent data on SNe projects. They
start the analysis with the data from Perlmutter et al. (1999), which is one of the important data
sets of the rst generation of SN Ia cosmology programs (Perlmutter et al. 1998; Garnavich
et al. 1998; Riess et al. 1998; Schmidt et al. 1998; Perlmutter et al. 1999). They particularly
focus on the sample of 54 SNe from their primary t C. Results from this tting procedure
are shown in Table 4.2. The CDM, as well as the models with the constant w have a good
t. For example the concordance model (at CDM) has a
2
= 57.7 at 52 dof with the
probability P = 27.3%, representing a good t. However, we also note that the data favour a
spherical universe. The Einstein de Sitter (EdS) model (
m
= 1, = 0) has a bad t with
P = 0.06%.
Table 4.2: Fits of different cosmologies to available data sets [24].
Models Constraint
m

or
X
w
2
dof P
54 SNe from Perlmutter et al. (1999)
CDM
m
+

= 1 0.28 0.08 1
m
-1 57.7 52 0.273
CDM none 0.79 0.47 1.40 0.65 -1 56.9 51 0.266
const w
m
+
X
= 1 0.48 0.15 1
m
2.10 1.83 57.2 51 0.257
EdS
m
= 1, = 0 1 1
m
92.9 53 0.0006
Gold sample of 157 SNe from Riess et al. (2004)
CDM
m
+

= 1 0.31 0.04 1
m
-1 177.1 155 0.108
CDM none 0.46 0.10 0.98 0.19 -1 175.0 154 0.118
const w
m
+
X
= 1 0.49 0.06 1
m
2.33 1.07 173.7 154 0.132
EdS
m
= 1, = 0 1 1
m
324.7 156 10
13
164 SNe from Gold + ESSENCE (Krisciunas et al. 2005)
CDM
m
+

= 1 0.30 0.04 1
m
-1 190.3 162 0.064
CDM none 0.48 0.10 1.04 0.18 -1 187.2 161 0.077
const w
m
+
X
= 1 0.50 0.06 1
m
2.69 1.30 185.5 161 0.091
EdS
m
= 1, = 0 1 1
m
348.2 163 10
15
168 SNe from Gold + ESSENCE + 4 SNe from Clocchiatti et al. (2005)
CDM
m
+

= 1 0.29 0.04 1
m
-1 200.79 166 0.034
CDM none 0.51 0.09 1.12 0.16 -1 195.8 165 0.051
const w
m
+
X
= 1 0.50 0.04 1
m
3.18 1.46 193.4 165 0.065
EdS
m
= 1, = 0 1 1
m
367.5 167 10
17
The existing data points coming froma wide range of different observations were compiled
by Tonry et al. [23]. With many new important additions towards higher redshifts, a rened
sample of 194 SNe was presented by Barris et al. (2004). However, the data we are going
to consider next, is the Gold sample of Riess et al. [18], which is a more reliable set of
data with reduced calibration errors arising from the systematics. It contains 143 points from
the previously published data, plus 14 new points with z > 1 discovered with the Hubble
Space Telescope (HST). While compiling this sample, various points from the previously
4.5 Measuring the Acceleration of the Present Universe 143
published data were discarded where the classication of the supernova was not certain or
the photometry was incomplete. The t-results of this data show that although the ts to
the CDM and the constant w-models are reasonable, they are deteriorated considerably; the
probabilities P, which represent the goodness-of-t, have reduced to less than half of the
corresponding probabilities obtained in the case of the Perlmutter et al. data.
We now consider the rst results of the ESSENCE project (Krisciunas et al. 2005, made
public in August 2005) under which 9 SNe with redshift in the range 0.5 - 0.8 were discovered
jointly with HST and Cerro Tololo 4-m telescope. In order to minimize the systematic errors,
all the ground-based photometry was obtained with the same telescope and instrument. We
consider 7 SNe of this project which have unambiguous redshift and denite classication,
and add them to the gold sample resulting in a reliable sample of 164 SNe. The Table 1
shows that the ts to different cosmologies have further worsened considerably and do not
represent good t, in any case. Increasing the number of tted parameters (for example, in the
models with a constant

) improves the t marginally only.


Next we consider the recent discovery of 5 SNe at redshift z 0.5 by the High-z Super-
nova Search Team (Clocchiatti et al. 2005, results made public in October 2005). We consider
4 SNe from this sample for which distances estimated from the MLCS2k2 (Multi-colour Light
Curve Shape) method are available, so that we can include them in the previous sample of
gold + ESSENCE which also use the MLCS2k2 method to determine the distance moduli.
The t-results of the resulting sample of 168 SNe are very disappointing. The quality of the
ts to different models has deteriorated to such extent that the concordance model can be re-
jected at 96.6 % condence level! This is an alarming situation. Other models have marginally
similar t and increasing the number of tted parameters does not help signicantly. Models
with variable

(t) do not help either. For example, if we consider w(z) = w


0
+w
1
z/(1+z)
with w
0
, w
1
as constants, we obtain
m
= 0.42,
X
= 0.42, w
0
= 4.95 and w
1
= 2.83 as
the best-tting solution with
2
= 193.07 at 163 dof and P = 5.4%.
Finally, the author also considers the recently published (made public in October 2005)
rst year-data of the planned ve-year SuperNova Legacy Survey (SNLS) (Astier et al. [1]).
In this survey the authors claim to have achieved high precision from improved statistics and
better control of systematics by using the multi-band rolling search technique and a single
imaging instrument to observe the same elds. Their data set includes 71 high redshift SNe
Ia in the redshift range 0.2 - 1 from the SNLS, together with 44 low redshift SNe Ia compiled
from the literature but analyzed in the same manner as the high-z sample. As the SNLS data
have been analyzed differently, the tting procedure followed by the authors is also different.
In order to calculate
2
, they use

2
=
N

j=1
_
{(z
j
; M,
i
)
obs,j
}
2

obs,j
+
2
int
_
, (4.211)
where
int
, is the (unknown) intrinsic dispersion of the SN absolute magnitude which, unlike
the other data sets, is not included in the

o
bs
; rather it has been used as an adjustable free
parameter to obtain
2
/dof = 1. We shall return to this issue later for our comments. First
we want to verify if we can reproduce the results of Astier et al. from equation (4.211) by
using their calculated
o
bs (given in columns 7 of their Tables 8 and 9) instead of using their
stretch and color parameters, which do not seem necessary once we have
o
bs. We nd
that by xing
int
= 0.13, we get
m
= 1

= 0.26, M = 43.16 with


2
/dof = 1.00;
and
m
= 0.31,

= 0.81, M= 43.15 with


2
/dof = 1.01. This is exactly what Astier et al
have obtained.
144 4 The Universe with Matter and Dark Energy
However, the introduction of
int
in equation (4.211) is justied only when we use in-
dependent measurement uncertainties
int,j
on the parameter, instead of using it as a free
parameter. The latter case is just equivalent to increasing the error bars suitably in order to
have a desired t. In this way one can t any model to the data. For example, the EdS model
can also have an excellent t by considering
int
= 0.258: M= 43.46 with
2
/dof = 1.00. It
should also be noted that this approach (which is equivalent to assuming that the data have a
good t to the model) prohibits an independent assessment of the goodness-of-t-probability
P, in the absence of which the estimated parameters do not have much signicance. However,
one can still compare the goodness-of-t of different models. For example, with
int
= 0.13,
the EdS model has a worse t than the CDM model, giving
2
/dof = 2.7. As the SNLS data
have been analyzed in a different way than the other available data, it does not make sense to
add this data to the earlier samples for a joint analysis.
The Highz Supernovae Search team has recently discovered 8 new SNe in the redshift
interval z = 0.3 1.2 [23]. From these observations they derive a SN Ia rate of (1.4 0.5)
10
4
h
3
Mpc
1
yr
1
at a mean redshift of 0.5. Including the constraint of a at Universe,
they nd
M
= 0.28 0.05.
Nesseris et al. [13] have performed a similar comparative analysis with three SNe datasets:
full Gold sample [18], a truncated Gold sample, and the SNLS dataset consiting of 115 SNe
[1].
GammaRay Burster as Standard Candles
One of the few ways to measure the properties of Dark Energy is to extend the Hubble daigram
(HD) to higher redshifts with Gamma-Ray Bursts (GRBs). GRBs have at least ve properties
(their spectral lag, variability, spectral peak photon energy, time of the jet break, and the
minimum rise time) which have correlations to the luminosity of varying quality. In Schaefer
2006, they construct a GRB HD with 69 GRBs over a redshift range of 0.17 to > 6, with half
the bursts having a redshift larger than 1.7. This paper uses over 3.6 times as many GRBs and
12.7 times as many luminosity indicators as any previous GRB HD work. The GRB HD is
well-behaved and nicely delineates the shape of the HD. The reduced chi-square for the t to
the concordance model is 1.05 and the RMS scatter about the concordance model is 0.65 mag.
The plot 4.24 shows the difference in distance modulus between the empty model and the
supernova and the GRB binned data. It looks a bit inconsistent at redshifts near 0.5 but the
residuals from the ts are not much bigger than the stated errors. When tting to both the SNe
and GRB datasets, there should be two free parameters for Hubble constant changes, one for
each dataset. These free parameters can be thought of as adjustments to the overall luminosity
calibration of SNe and GRBs respectively.
The Future
How far can we push the SN measurements ? Finding more and more SNe allows us to beat
down statistical errors to arbitrarily small levels but, ultimately, systematic effects will limit
the precision to which SNIa magnitudes can be applied to measure distances (Fig. 4.25).
Our best estimate is that it will be possible to control systematic effects from groundbased
experiments to a level of 0.03 mag. Carefully controlled ground-based experiments on 200
SNe will reach this statistical uncertainty in z = 0.1 redshift bins in the range z = 0.3 0.7,
and is achievable within ve years. A comparable quality low redshift sample, with 300 SNe
in z = 0.02 0.08, will also be achievable in that time frame.
4.5 Measuring the Acceleration of the Present Universe 145
Figure 4.24: The difference in distance modulus between the empty model and the supernova
and the GRB binned data. The GRB data are from Schaefer (2006). [Plot: Ned Wright 2008]
Figure 4.25: SNAP constraints on dark energy models. The current set of 18 + 42 SNe is
supplemented by a set of SNAP SNe (each SNAP point represents data of 50 supernovae).
The SuperNova/Acceleration Probe (SNAP) collaboration has proposed to launch a dedi-
146 4 The Universe with Matter and Dark Energy
cated cosmology satellite - the ultimate SNIa experiment. This satellite will, if funded, scan
many square degrees of sky, discovering well over a thousand SNIa per year and obtain their
spectra and light curves out to z = 1.7. Besides the large numbers of objects and their ex-
tended redshift range, space-based observations will also provide the opportunity to control
many systematic effects better than from the ground. Fig. 4.26 shows the expected precision
in the SNAP and groundbased experiments for measuring w, assuming a at Universe. Per-
haps the most important advance will be the rst studies of the time variation of the equation
of state w (see the right panel of Fig. 4.26).
Figure 4.26: Future expected constraints on dark energy: Left panel: Estimated 68% condence
regions for a constant equation of state parameter for the dark energy, w, versus mass density,
for a ground-based study with 200 SNe between z = 0.3 0.7 (open contours), and for the
satellite-based SNAP experiment with 2,000 SNe between z = 0.31.7 (lled contours). Both
experiments are assumed to also use 300 SNe between z = 0.02 0.08. A at cosmology
is assumed (based on Cosmic Microwave Background (CMB) constraints) and the inner (solid
line) contours for each experiment include tight constraints (from large scale structure surveys)
on
M
, at the 0.03 level. Such ground-based studies will test the hypothesis that the dark
energy is in the form of a cosmological constant, for which w = 1 at all times. Middle
panel: The same condence regions for the same experiments not assuming the equation of
state parameter, w, to be constant, but instead marginalizing over w, where w(z) = w0 + wz.
(Weller and Albrecht [25] recommend this parameterization of w(z) over the others that have
been proposed to characterize well the current range of dark energy models.) Note that these
planned ground-based studies will yield impressive constraints on the value of w today, w0,
even without assuming constant w. In fact, these constraints are comparable to the current
measurements of w assuming it is constant. Right panel: Estimated 68% condence regions
of the rst derivative of the equation of state, w

, versus its value today, w


0
, for the same
experiments.
With rapidly improving CMBdata frominterferometers, the satellites Microwave Anisotropy
Probe (MAP) and Planck, and balloon-based instrumentation planned for the next several
4.6 Angular Width in FRW Models 147
years, CMB measurements promise dramatic improvements in precision on many of the cos-
mological parameters. However, the CMB measurements are relatively insensitive to the dark
energy and the epoch of cosmic acceleration. SNIa are currently the only way to directly
study this acceleration epoch with sufcient precision (and control on systematic uncertain-
ties) that we can investigate the properties of the dark energy, and any time dependence in
these properties. This ambitious goal will require complementary and supporting measure-
ments of, for example,
M
from CMB, weak lensing, and large scale structure. The SN mea-
surements will also provide a test of the cosmological results independent from these other
techniques, which have their own systematic errors. Moving forward simultaneously on these
experimental fronts offers the plausible and exciting possibility of achieving a comprehensive
measurement of the fundamental properties of our Universe.
4.6 Angular Width in FRW Models
We already have derived a general expression for the angular width of an object in an
expanding Universe, based on the angular distance which is related to the luminosity distance
=
D(1 +z)
2
d
L
. (4.212)
The redshift factor in the nominator cancels to a large extent the increase in the luminosity
distance with redshift. For a Universe dominated by vacuum energy, the apparent angular
width reaches a constant value which is just given by D/R
H
(Fig. 4.27). For all other
models, the angular width attains a minimum at some redshift around z 1. For matter
dominated models with

= 0 (SCDM), q
0
= 1/2 the minimum can easily be determined
=
DH
0
2c
(1 +z)
3/2

1 +z 1
. (4.213)
We attain a minimum for
z
min
= 1.25 (4.214)
and the corresponding minimal angular width is given by
4

min
= 3.375
DH
0
c
= 0.12 (2h) mas
_
D
pc
_
. (4.215)
In the SCDM model, a galaxy with dimension of 30 kpc would have an angular diameter
of 3.6 arcsec at redshift z = 1.2. It would be nicely resolvable with presentday telescopes.
At higher redshifts, the angular diameter would even increase. In fact, on deep images of the
Sky (such as HubbleDeep eld or the FORSDeep eld) galaxies appear resolved and have
typical extensions of about one to a few arcseconds. This focussing property of the expanding
Universe is essential for observations of highredshift objects. The jets of quasars remain
more or less of constant angular extension beyond redshift one. It is however very difcult
to nd a cosmic standard rod which could be used to measure the cosmological parameters
(extension of galaxy clusters or the length of compact jets in quasars).
4
1 mas = 1 milliarcsecond
148 4 The Universe with Matter and Dark Energy
z
10
-1
1 10
D
/
R
H
u
b
b
l
e
1
10
Angular Width Inflationary Universe
Figure 4.27: Apparent angular width as a function of redshift in the at inationary Universe.
The decay with 1/z corresponds to the classical redshift behaviour without expansion, the blue
solid line to a matterdominated at model with
M
= 1 (SCDM), the dotted line to a matter
dominated open model for
M
= 0.3 (OCDM), the star symbols refer to a CDMmodel with

M
= 0.3,

= 0.7, the lowest curve to a vacuum dominated de Sitter Universe with

= 1
(in all models: H
0
= 70 km/s/Mpc).
Angular Width of Structure at Recombination
For the interpretation of the anisotropies in the CMBR it is important to know what the intrin-
sic scale is e.g. for an angular width of one degree. As we see from Fig. 4.28, the SCDM
model provides an upper limit for the angular width
(rec) <
DH
0
c
z
rec
2
. (4.216)
A reasonable value is (rec) 300 DH
0
/c. A scale of one degree corresponds therefore
to a structure D 0.2 Mpc at recombination, i.e. a structure of 200 Mpc today. Typical
temperature uctuations have been measured to occur on angular scales of about half a degree
(WMAP, Fig. 2.9), corresponding to a scale of roughly 100 Mpc in the present Universe. This
is in fact the mean distance between large clusters which form the corners of huge laments
(Fig. 4.29). There is no structure beyond this scale, i.e. the homogeneity is indeed satised on
those scales. This is in agreement with the ndings that the CMBR spectrum is at on scales
larger than about one degree.
The Millennium Simulation (Fig. 4.29) employed more than 10 billion particles of mat-
ter to trace the evolution of the matter distribution in a cubic region of the Universe over 2
billion light-years on a side.
5
It kept the principal supercomputer at the Max Planck Societys
5
The Virgo Consortium is an international grouping of scientists carrying out supercomputer simulations of the
4.7 Redshift Distribution of Cosmological Objects 149
10
-1
1 10 10
2
10
3
1
10
10
2
Angular Width Recombination
Vacuum
z
SCDM
OCDM
Figure 4.28: Apparent angular width in units of D/R
H
as a function of redshift at the recombi-
nation era. The highest angular widths are obtained for classical SCDM models, while vacuum
energy reduces the angular width (starry line: CDM with
M
= 0.3). In a CDM model we
nd for the angular width at recombination (Rec) 300 D/R
H
.
Supercomputing Centre in Garching, Germany occupied for more than a month. By applying
sophisticated modelling techniques to the 25 Terabytes of stored output, Virgo scientists are
able to recreate evolutionary histories for the approximately 20 million galaxies which popu-
late this enormous volume and for the supermassive black holes occasionally seen as quasars
at their hearts.
4.7 Redshift Distribution of Cosmological Objects
We assume that there are n(t
1
; L) dL galaxies at time t
1
having absolute luminosity between
L and L +dL. The number of such objects is then given by
dN =
R(t
1
) dr
1
_
1 kr
2
1
R
2
(t
1
)r
2
1
n(t
1
; L) dLd. (4.217)
Since
dr
1
=
_
1 kr
2
1
c dt
1
/R(t
1
) (4.218)
formation of galaxies, galaxy clusters, large-scale structure, and of the evolution of the intergalactic medium. Al-
though most of the members are based in Britain and at the MPA in Germany, there are important nodes in Japan,
Canada and the United States. The primary platforms used by the consortium are the IBM supercomputer at the
Computing Centre of the Max Planck Society in Garching and the Sun Microsystems Cosmology Machine at the
Institute for Computational Cosmology of Durham University.
150 4 The Universe with Matter and Dark Energy
Figure 4.29: The distribution of dark matter in the Universe on different scales. The background
picture shows a cut through the Millennium Simulation with a total extension of more than 9
billion lightyears on a side. On such huge scales, the universe appears nearly homogeneous, but
the series of enlargements overlayed show a complex cosmic web of dark matter up to scales of
order 100 Mpc. This large-scale structure consists of laments which surround large voids
and cross in massive halos of matter. The largest of these halos are rich clusters of galaxies,
containing more than one thousand galaxies which are still resolved as halo substructure in the
simulation. [22]
4.7 Redshift Distribution of Cosmological Objects 151
we nd
dN = R
2
(t
1
)r
2
1
n(t
1
; L) c|dt
1
| dLd. (4.219)
From the redshift condition
dt =
dz
(1 +z) E(z)
(4.220)
the number is given by
dN = R
H
R
2
(t
0
)r
2
1
n(t
1
; L)
dz dLd
(1 +z)
3
E(z)
. (4.221)
This may be expressed in terms of the luminosity distance
dN = R
H
d
2
L
(z)
(1 +z)
5
E(z)
n(t
1
; L) dz dLd. (4.222)
For a classical Friedmann Universe,

= 0, we recover the old expression


dN = R
H
d
2
L
(z)
(1 +z)
6

1 +
M
z
n(t
1
; L) dz dLd. (4.223)
This is equivalent to the expression derived in Sect. 3.6 for the number counts of galaxies per
unit solid angle and per redshift interval
Redshift
10
-1
1 10
(
1
/
V
_
H
)

d
V

/

d
z

d
O
10
-3
10
-2
OCDM
S
C
D
M
L
C
D
M
Comoving Volume Inflationary Universe
Figure 4.30: The comoving cosmic volume per redshift interval and per steradian as a function of
redshift for various FRW models (in units of the Hubble volume V
H
= R
3
H
). The highest volume is
achieved for models including dark energy, while in an inationary Universe with vanishing dark energy,
the lowest volume is obtained.
dN
dz d
= n(z) d
H
(z)
d
2
L
(z)
(1 +z)
5
(4.224)
152 4 The Universe with Matter and Dark Energy
d
H
(z) = c/H(z) is the Hubbleradius at redshift z and d
L
(z) the luminosity distance. For
objects which are conserved per proper volume, n remains constant as a function of redshift,
or
n(z; L) = n(0; L) (1 +z)
3
, (4.225)
we obtain for the total distribution
dN(z) = R
H
_

0
dL
d
2
L
(z)
(1 +z)
2
E(z)
n(0; L) dz d. (4.226)
For a class of objects with timeindependent luminosity (Supernovae e.g.) we would have a
distribution of the kind
dN(z) = R
H
d
2
L
(z)
(1 +z)
2
E(z)
n(0) dz d. (4.227)
In distinction to the above proper volume, the comoving volume dened as
dV
C
= d
H
(z) d
2
C
(z) dz d = d
H
(z)
d
2
L
(1 +z)
2
dz d (4.228)
is often used. Here, d
C
(z) = R
0
r
em
is the comoving distance. When compared to a non
expanding Universe, the proper volume attains a maximum around redshift one (Fig. 4.30).
The location of the maximum depends on the particular FRWmodel.
In Fig. 4.31 we show the behaviour of the differential number counts according to relation
(4.227). In all FRWmodels, the number counts reach a maximum around redshift two and
then slowly decline towards higher redshifts. In uxlimited samples, this decline would
occur much more rapidly, since at high redshifts only the brightest sources can be observed.
If we were observing galaxies with sufcient sensitivity, we should see this maximum around
redshift two. This behaviour is typically seen in the redshift distribution of Quasars. It is,
however, quite interesting that the redshift distribution of Quasars attains its maximum around
redshift two.
How much of the Sky is covered by galaxies ? Another interesting question is the proba-
bility of a given line of sight to intersect a galaxy with redshift in the interval [z, z +dz]. This
is given by
dP
G
= r
2
G
ndl (4.229)
resulting in the expression
dP
G
dz
= r
2
G
n(z)
d
H
(z)
1 +z
. (4.230)
If galaxies are neither created nor destroyed, n(z) = n
0
(1 +z)
3
, showing that
dP
G
dz
= r
2
G
R
H
n
0
_
(d
H
(z)/R
H
) (1 +z)
2
_
. (4.231)
4.8 The Cosmological Fundamental Plane 153
Redshift
0 1 2 3 4 5
(
1
/
n
V
_
H
)

d
N

/

d
z

d
O
0
0.1
0.2
0.3
0.4
0.5
OCDM
SCDM
L
C
D
M
Number Counts Inflationary Universe
Figure 4.31: The number counts of galaxies per redshift interval and per steradian as a function of
redshift for various FRW models (in units of the Hubble volume number nV
H
). In all models, the
observed number counts reach a maximumaround redshift two and then decline towards higher redshifts.
The total optical depth for intersection of objects upto some redshift z is then given by the
integral

G
(z) =
_
z
0
dP
G
= r
2
G
R
H
n
0
_
z
0
dx(d
H
(z)/R
H
)(1 +x)
2
. (4.232)
This integral can be done e.g. in a classical inationary FRWmodel with
M
= 1, where
d
H
(z) = R
H
/(1 +z)
3/2
,

G
(z) =
2
3
n
0
r
2
G
R
H
(1 +z)
3/2
. (4.233)
For bright galaxies, n
0
0.02 h
3
Mpc
3
and r
G
10/h kpc, we get
G
(z) 0.01 (1+z)
3/2
.
At redshift one, about 4% of the Sky is covered by galaxies. The optical depth would reach
unity at a redshift of z 20, but galaxies are formed at redshifts lower than this value.
4.8 The Cosmological Fundamental Plane
Due to the restriction
k
+
M
+

= 1, given by the Friedmann equation, the present state of


the Universe is determined by a point in the (
M
,

)plane. This plane is nowadays called


Fundamental Plane of Cosmology (Fig. 4.32). In particular, the transition from accelerated
to decelerated models is given by the line q
0
= 0. Three types of data are available to constrain
the location of the present Universe
Supernovae data, as discussed in the previous section;
154 4 The Universe with Matter and Dark Energy
CMBR anisotropies (WMAP and other experiments, see Sect. 8);
mass determinations in clusters of galaxies (Chandra and XMMNewton, Sect. 2.4).
All these data are compatible with the assumption of a nonvanishing vacuum energy. The
bestbuy model of the year 2005 (called concordance model) is given by the following pa-
rameters
H
0
= (71 5) km/s/Mpc

k
= 0.02 0.02 (the Universe is at)

M
= (0.135 0.08)/h
2
= 0.27 0.04

= 0.73 0.04

B
= (0.0224 0.0009)/h
2
= 0.044 0.004.
Dark Energy is compatible with vacuum energy, w
0
= 1.
It is the rst time since 1929 that realistic error bars are given for these parameters. The
Universe is completely dominated by Dark Energy and Dark Matter.
4.9 On the Origin of the Dark Energy
Not much is known about dark matter, and even less about dark energy. Cosmologists have
taken to discussing the enigmatic properties of the dark energy with the use of a new pa-
rameter, w, which is the ratio of its average pressure to energy density. The degree of this
runaway expansion impulse is expressed by w. What is the nature of dark energy and how
does it overcome the attractive pull of gravitation in order to speed up the cosmic expansion,
and what is the proper value of w? In the best known model, the cosmological constant in
Einsteins famous equations of general relativity corresponds to energy and pressure of the
universal quantum vacuum, and is constant in space and time. Here the value of w is -1. In a
second popular model, the quintessence model, the dark energy is associated with a universal
quantum eld relaxing towards some nal state. Here the energy density and pressure of the
dark energy are slowly decreasing with time, and the value of w is somewhere between -1/3
and -1 (w must be smaller than -1/3 in order for cosmic acceleration to occur).
A new cosmic doomsday scenario takes the present acceleration of the expansion of the
universe to new extremes. Dartmouth physicist Robert Caldwell and his colleagues Marc
Kamionkowski and Nevin Weinberg at Caltech have determined that if the supposed dark en-
ergy responsible for the acceleration is potent enough not only will the space between galaxies
continue to increase but that the galaxies themselves will y apart as will, at successive times
stars, planets, and even atoms and nuclei.
In Caldwells phantom energy model, there is no stable vacuum quantum state and the
energy density and the expansionary pressure exerted on the universe seems to increase even
as the spacetime itself expands (with ordinary gases, pressure falls with expansion). In this
scenario w is less than -1. The implications of this new type of cosmology are that bound
systems should in the course of time be ripped up. For example, at a w value of -1.5 the
universe would last for 35 billion years before being ripped apart. About 60 million years
before the end, the Milky Way would be torn apart. About 3 months before the end the so-
lar system would become undone. About 30 minutes before that the Earth would explode.
4.9 On the Origin of the Dark Energy 155
Figure 4.32: The Fundamental Plane of Cosmology with the restrictions from SNe, CMBR
anisotropies and dark matter in clusters. All these data are compatible with a at Universe
dominated by dark energy. In the future, the question whether the Universe is really at can be
attacked (SNAP e.g.).
And about 10
19
seconds before the ultimate moment of doom, atoms would be pulled apart.
Caldwell suggests that deciding between this model and the others might be possible in com-
ing years with much better data coming from microwave background, supernovae, and galaxy
measurements [3].
156 4 The Universe with Matter and Dark Energy
4.9.1 The Vacuum is not Empty
According to quantum mechanics, the vacuum is not empty, but teeming with virtual particles
that constantly wink in and out of existence. One strange consequence of this sea of activity is
the Casimir effect: Two at metal surfaces automatically attract one another if they get close
enough. The Casimir force is so weak that it has rarely been detected at all, but now a team
reports in the 23 November PRL that they have made the most precise measurement ever of
the phenomenon. They claim that their technique, using an atomic force microscope, has the
capacity to test the strangest aspects of the Casimir effect, ones that have never before been
tested.
The simplest explanation of the Casimir effect is that the two metal plates attract because
their reective surfaces exclude virtual photons of wavelengths longer than the separation
distance. This reduces the energy density between the plates compared with that outside, and
like external air pressure tending to collapse a slightly evacuated vesselthe Casimir force
pulls the plates toward one another. But the most puzzling aspect of the theory is that the
force depends on geometry: If the plates are replaced by hemispherical shells, the force is
repulsive. Spherical surfaces somehow enhanced the number of virtual photons. There is no
simple or intuitive way to tell which way the force will go before carrying out the complicated
calculations.
Since the discovery of the theory by Casimir 50 years ago, there have been only two
previous documented detections of the effect. One was in 1958 and had 100% uncertainty,
and the second was in 1997, when the theory was veried to within 5%. Umar Mohideen
and Anushree Roy, of the University of California at Riverside, claim their results verify the
theory to within 1% [Phys. Rev. Lett. 81, 4549 (1998)].
Technically, quantum elds have an innite number of possible energy states, all of which
should contribute virtual particles to the vacuum. Theorists normally assume that the num-
ber of states is actually nite because they cant exceed the so-called Planck energy, which
corresponds to the smallest distance quantum mechanics allows for.
4.9.2 Quintessence
Todays cosmologists nd Lambda to be just as objectionable, but for a different reason. All
quantum elds possess a nite amount of zero-point vacuum energy as a result of the uncer-
tainty principle. A naive estimate of the zero-point energy predicts a vacuum energy density
that is 120 orders of magnitude greater than the energy density of all the other matter in the
universe. If the vacuum energy density really is so enormous, it would cause an exponentially
rapid expansion of the universe that would rip apart all the electrostatic and nuclear bonds that
hold atoms and molecules together. There would be no galaxies, stars or life. Since we cannot
ignore quantum mechanics, some other mechanism must nullify this vacuum energy. One of
the major goals of unied theories of gravity has been to explain why the vacuum energy is
zero.
Einsteins blunder has been resurrected as a possible solution to the dark-energy problem.
Maybe there is a miraculous cancellation mechanism, but perhaps it is slightly imperfect.
Instead of making Lambda exactly zero, the mechanism only cancels to 120 decimal places.
Then, vacuum energy would comprise the missing two-thirds of the critical density. The
requirements seem bizarre, though. Some constant that is naturally enormous must be cut
down by 120 orders of magnitude, but with such precision that today it has just the right value
to account for the missing energy.
4.9 On the Origin of the Dark Energy 157
Extrapolating back in time to the early universe, the story seems even more bizarre. When
the volume of the universe was 100 orders of magnitude smaller, say, the mass density was
100 orders of magnitude greater, but the vacuum energy density had to have the same value as
today. In other words, the vacuum energy density remained constant as the universe expanded,
but the total vacuum energy increased as the volume of space increased. This extra energy
came from the gravitational potential energy of the universe. Whatever physical processes
created the initial energy in the universe had to arrange for an exponentially large difference
between the two forms of energy, but somehow this difference had to have exactly the right
value for the vacuum energy to become important 15 billion years later.
It would seem more natural for the dark energy to start with an energy density similar
to the density of matter and radiation in the early universe (Fig. 4.33). The dark energy and
matter density could both then decrease at similar rates as the universe expanded, with the dark
energy density overtaking the matter density only after structure has formed in the universe.
However, if the dark energy density has been changing, it cannot consist of vacuum energy.
Therefore, the concept of quintessence was introduced to overcome this problem in 1998.
Quintessence is a dynamic, time-evolving and spatially dependent form of energy with
negative pressure sufcient to drive the accelerating expansion. Whereas the cosmological
constant is a very specic form of energy - vacuum energy - quintessence encompasses a wide
class of possibilities.
The simplest model proposes that the quintessence is a quantum eld with a very long
wavelength, approximately the size of the observable universe. Some examples had been
explored almost a decade earlier by Bharat Ratra and James Peebles at Princeton University,
and by Wetterich (Heidelberg). A particle is usually thought of as a bundle of oscillations in
a quantum eld, but since this bundle is much larger than any conventional length scale, the
particle description is impractical.
The energy is composed of kinetic energy, which depends on the rate of oscillations in the
eld strength, and potential energy, which depends on the interaction of the eld with itself and
matter. The pressure is determined by the difference between the kinetic and potential energy,
with kinetic energy contributing positively to the pressure. However, since the oscillation has
an extremely long wavelength and period - essentially the size and age of the universe - its
kinetic energy is negligible. The behaviour of the quintessence eld is therefore dominated by
how it interacts with itself. Much like a stretched spring, this self-interaction potential leads
to negative pressure.
Within certain models that seek to unify the four fundamental forces of nature there exist
elds, called tracker elds, that can make quintessence behave in this way (see Zlatev et al.
in further reading). First, the dark energy density in the early universe can be comparable with
the matter density. The model is insensitive to the precise initial value because the dynamical
equations that determine the time evolution of the tracker eld have solutions that cause the
energy to follow the same evolution, independent of initial conditions (similar to the classical
attractor solutions found in conventional nonlinear dynamics). In particular, the energy density
tracks the radiation and matter density (see Figure 4.33)). For most of the history of the
universe, the quintessence occupies a very small fraction of the critical density, but the fraction
grows slowly until it catches up with and ultimately overtakes the matter density.
It seems natural to ask if there are any direct gravitational interactions between ordinary
matter and dark energy. If the dark energy is vacuum energy, then the two do not interact
because vacuum energy is inert and unchanging. But if the dark energy is quintessence, they
can interact under certain conditions. Ordinary particles are physically very small compared
158 4 The Universe with Matter and Dark Energy
Figure 4.33: This gure illustrates the different scaling of energy densities of radiation (dotted
curve), matter (dashed curve), a cosmological constant and two quintessence models. The solid
green line is an early dark energy model where the amount of dark energy is non-negligible at
early times. The dashed green line depicts a phantom dark energy model. a is the scale factor,
normalized such that a = 1 today. The energy scale at the right side of the gure is conveniently
chosen as Gev per cubic meter: one hydrogen atom corresponds to one Gev and hence the mean
density of our Universe today corresponds to a few hydrogen atoms per cubic meter.
with the Compton wavelength,
c
= /mc, of the quintessence particles, so individual par-
ticles have a completely negligible effect on quintessence (and vice versa). However, very
large clumps of ordinary matter - spread out over a region comparable with the Compton
wavelength - can interact gravitationally with quintessence and create inhomogeneities in its
distribution that may produce detectable signals in the cosmic microwave background.
What cosmologists nd most difcult to explain is why the acceleration should begin at
this particular moment in cosmic history. Is it a coincidence that, just when thinking beings
have evolved, the universe suddenly shifts into overdrive? The situation is peculiar because
the energy associated with the cosmological constant or quintessence is very tiny, less than a
millielectron-volt. If new ultra-low-energy physics is responsible, it should have already been
observed in other experiments.
Perhaps a more satisfying possibility is that the acceleration is triggered by natural events
in the recent history of the universe. According to the big-bang model, the energy density
in the universe was predominantly in the form of hot, relativistic particles until the universe
was a few tens of thousands of years old. At that time, the universe had cooled enough that
the mass energy of non-relativistic particles became more important than both their kinetic
energy and the energy of radiation, resulting in a change in the cosmic expansion rate. This
marked the beginning of the matter-dominated epoch. Only then could gravity begin to
4.9 On the Origin of the Dark Energy 159
clump matter together to form stars, galaxies and large-scale structure. Is it possible that this
transition triggered the onset of quintessence?
Quintessence leaves its mark on the universe in several ways, so experimenters have a
number of methods they can use to test for this exotic form of energy. The acceleration effect
of a dark-energy component depends on the ratio of its pressure to its energy density. More
negative values of this ratio, w, lead to greater acceleration. Quintessence and vacuum energy
have different values of w, so more precise measurements of supernovae over a longer span of
distances may be able to separate these two possibilities.
This challenge is the motivation for two proposals - the Earth-based Large-Aperture Syn-
optic Survey Telescope (LSST) and the space-based Supernova Acceleration Project (SNAP)
- that will monitor the sky for supernovae and other time-varying astrophysical phenomena.
The proposers of these projects are currently seeking funding.
Cosmic acceleration also affects the number of galaxies to be found as one explores deeper
and deeper into space. With appropriate corrections for evolution and other effects, the av-
erage density of galaxies is uniform throughout space. Consequently, for a xed range of
distances, one should nd the same number of galaxies nearby and far away. But cosmolo-
gists measure the redshift of distant galaxies, not their distance. The conversion from redshift
to distance follows a simple linear relation (the Hubble law) if the distances are small, but a
nonlinear relation depending on the acceleration of the universe if the distances are large. The
nonlinear relation will cause the number of galaxies found for a xed range of redshifts to
change systematically as one probes deeper into space. The Deep Extragalactic Evolutionary
Probe (DEEP), an advanced spectrograph on the Keck II telescope in Hawaii, is poised to test
this prediction with an accuracy that may be sufcient to distinguish between quintessence
and a cosmological constant.
Quintessence should also have an effect on the cosmic microwave background because
differences in the acceleration rate will produce small differences in the angular size of hot
and cold spots. Moreover, unlike a cosmological constant, quintessence is not spatially ho-
mogeneous. Small variations in the amount of quintessence across the sky should be seen as
ripples in the microwave background temperature. Measurements by the WMAP and Planck
satellites may be able to detect these effects, which will be at the level of a few per cent,
although it will be difcult to separate them from other effects.
In many cases, quintessence interacts with matter in a way that affects the forces between
particles. Then, if the quintessence eld is varying temporally or spatially, it will cause the
strengths of the forces between particles to change as well. Hence, ongoing tests for changes
in the values of the fundamental physical constants with time could be another source of
evidence for quintessence. It might be possible to search for such effects with astrophysical
observations (e.g. a variation of hyperne splitting with redshift) or in ultrahigh-precision
laserspectroscopy experiments.
4.9.3 Braneworld Models Higher Dimensions
Pretend you lived on your computer screen and could only move on that two dimensional
surface. The computer exists in three space dimensions but you can only move on a two
dimensional subspace made by the screen, so the spacetime that you experience would look
like three dimensions (two space plus time) rather than four. Thats sort of the idea in a
braneworld higher dimensional theory. Our observed four dimensional spacetime is like the
computer screen, a subspace of some bigger space that we cannot see because all matter and
160 4 The Universe with Matter and Dark Energy
forces are constrained to move (mainly) on our subspace, or brane (as in membrane). The total
space is called the bulk and the subspace or brane on which we would live is called the brane.
Gravity is the force that determines the shape of spacetime. Therefore, at least in principle,
gravity should propagate in all dimensions equally. That means we should be able to detect
large extra dimensions by looking for suspicious behavior in the gravitational force. But in
certain braneworld models, gravity can actually be conned or bound to our brane so that it
does not propagate very far in the bulk. That makes the extra dimensions harder to detect
using gravity.
Braneworld models have been extensively applied to cosmology, where they demonstrate
qualitatively new and very interesting properties (see [12, 19] for recent reviews). Theories
with the simplest generic action involving scalar-curvature terms both in the bulk and on the
brane are not only good in modeling cosmological dark energy but, in doing so, they also
exhibit some interesting specic features, for example, the possibility of superacceleration
(supernegative effective equation of state of dark energy w
e
1) and the possibility of
cosmological loitering even in a spatially at universe.
Figure 4.34: The Universe is considered to be a Brane in a 5dimensional spacetime, called bulk. Only
gravity can expand into the bulk, matter is conned to the Brane.
It was noted some time ago that the cosmological evolution in braneworld theory, from
the viewpoint of the Friedmann universe, proceeds with a time-dependent gravitational con-
stant. It turns out that, for a broad range of parameter values, the braneworld model behaves
exactly as a LCDM ( + Cold Dark Matter) universe with different values of the effective
cosmological density parameter
m
at different epochs. This allows the model to share many
of the attractive features of LCDM, but with the important new property that the cosmological
density parameter inferred from the observations of the large-sale structure and cosmic mi-
crowave background (CMB) and that determined from neoclassical cosmological tests such
as observations of supernovae (SN) can potentially have different values.
An important feature of this model is that, although it is very similar to LCDM at the
present epoch, its departure from concordance cosmology can be signicant at intermediate
redshifts, leading to new possibilities for the universe at the end of the dark ages which may
be worth exploring.
4.9 On the Origin of the Dark Energy 161
A Simple Model
We consider the simplest generic braneworld model with an action of the form
S = M
3
__
bulk
(R2
b
) 2
_
brane
K
_
+
_
brane
_
m
2
R 2
_
+
_
brane
L(h
ab
, ) .
(4.234)
Here, R is the scalar curvature of the metric g
AB
(A, B = 0, ..., 4) in the ve-dimensional
bulk, and R is the scalar curvature of the induced metric h
AB
= g
AB
n
A
n
B
on the brane,
where n
a
is the vector eld of the inner unit normal to the brane, which is assumed to be a
boundary of the bulk space. The quantity K = h
AB
K
AB
is the trace of the symmetric tensor
of extrinsic curvature K
AB
= h
C
A

C
n
B
of the brane. The symbol L(h
ab
, ) denotes the
Lagrangian density of the four-dimensional matter elds whose dynamics is restricted to
the brane so that they interact only with the induced metric h
ab
. All integrations over the bulk
and brane are taken with the corresponding natural volume elements. The symbols M and
m =
_
c/G denote the ve-dimensional and four-dimensional Planck masses, respectively,

b
is the bulk cosmological constant, and is the brane tension. In general, M m.
Action (4.234) leads to the Einstein equation with cosmological constant in the bulk:
G
AB
+
b
g
AB
= 0 , (4.235)
while the eld equation on the brane is
m
2
G
ab
+h
ab
=
ab
+M
3
(K
ab
h
ab
K) , (4.236)
where
ab
is the stressenergy tensor on the brane stemming from the last term in action
(4.234).
The cosmological evolution on the brane that follows from(4.235) and (4.236) is described
by the main equation [21, 4, 12]
H
2
+

a
2
=
+
3m
2
+
2

2
_
1

1 +
2
_
+
3m
2


b
6

C
a
4
_
_
, (4.237)
where C is the integration constant describing the so-called dark radiation and correspond-
ing to the black-hole mass of the Schwarzschild(anti)-de Sitter solution in the bulk, H a/a
is the Hubble parameter on the brane, and the term /a
2
corresponds to the spatial curvature
on the brane. The length scale is dened as
=
2m
2
M
3
= 2
_
m
M
_
2

P
(M) ,
P
(M) = /Mc . (4.238)
This length scale can be comparable to the Hubble radius c/H
0
, m = 1.3 10
19
GeV/c
2
and M > 1 GeV/c
2
. On short scalelengths, r , one recovers Einsteins gravity, on large
scalelengths, r , brane specic effects become important, leading to acceleration of the
Universe.
The signs in (4.237) correspond to two different branches of the braneworld solutions.
Models with the lower () sign were called Brane 1, and models with the upper (+) sign
were called Brane 2.
162 4 The Universe with Matter and Dark Energy
In what follows, we consider a spatially at universe ( = 0) without dark radiation
(C = 0). It is convenient to introduce the dimensionless cosmological parameters

m
=

0
3m
2
H
2
0
,

=

3m
2
H
2
0
,

=
1

2
H
2
0
,

b
=

b
6H
2
0
, (4.239)
where the subscript 0 refers to the current values of cosmological quantities. The cosmo-
logical equation with the energy density dominated by dust-like matter can now be written
in a transparent form:
H
2
(z)
H
2
0
=
m
(1+z)
3
+

+2

2
_

m
(1+z)
3
+

b
. (4.240)
The model satises the constraint equation

m
+

2
_

_
1 +

b
= 1 (4.241)
reducing the number of independent parameters. The sign choices in Eqs. (4.240) and
(4.241) always correspond to each other if 1+

b
>

. In the opposite case, 1+

b
<

,
both signs in (4.241) correspond to the lower sign in (4.240), and the option with both upper
signs in Eqs. (4.240) and (4.241) does not exist. The signs in (4.241) correspond to the two
possible ways of bounding the Schwarzschild(anti)-de Sitter bulk space by the brane [4].
In the normal case 1 +

b
>

, substituting

from (4.241) into (4.240), we get


H
2
(z)
H
2
0
=
m
(1+z)
3
+ 1
m
+ 2

2
_

_
1 +

b
2
_

m
(1+z)
3

m
+
_
_
1 +

b

_

_
2
. (4.242)
In the special case 1 +

b
<

, the Brane 1 model [the lower sign in (4.240)] corre-


sponds to both signs of (4.241), and the cosmological equation reads as follows:
H
2
(z)
H
2
0
=
m
(1+z)
3
+ 1
m
+ 2

2
_

_
1 +

b
2
_

m
(1+z)
3

m
+
_
_

_
1 +

b
_
2
. (4.243)
For sufciently high redshifts, the rst term on the right-hand side of (4.242) or (4.243)
dominates, and the model reproduces the matter-dominated Friedmann universe with the den-
sity parameter
m
. Now we note that, for the values of z and parameters

b
and

which
satisfy

m
(1+z)
3

_
_
1 +

b

_

_
2
, (4.244)
both Eqs. (4.242) and (4.243) can be well approximated as
H
2
(z)
H
2
0

m
(1+z)
3
+ 1
m

_
1 +

b
_

m
(1+z)
3

. (4.245)
4.9 On the Origin of the Dark Energy 163
We introduce the positive parameter by the equation
=
_
1 +

. (4.246)
Then, dening a new density parameter by the relation

LCDM
m
=

1

m
, (4.247)
we get
H
2
(z)
H
2
0

LCDM
m
(1+z)
3
+ 1
LCDM
m
, (4.248)
which is precisely the Hubble parameter describing a LCDMuniverse. [Note that the braneworld
parameters

and

b
have been effectively absorbed into a renormalization of the mat-
ter density
m

LCDM
m
, dened by (4.247). Since the value of the parameter
LCDM
m
should be positive by virtue of the cosmological considerations, in the case of the upper sign
in (4.245) and (4.247), we restrict ourselves to the parameter region > 1, i.e., we exclude
the special case of Brane 1 model with the upper sign in (4.241) from consideration.]
Thus, our braneworld displays the following remarkable behaviour which we refer to as
cosmic mimicry:
A Brane 1 model, which at high redshifts expands with density parameter
m
, at lower
redshifts masquerades as a LCDM universe with a smaller value of the density parame-
ter. In other words, at low redshifts, the Brane 1 universe expands as the LCDM model
(4.248) with
LCDM
m
<
m
[where
LCDM
m
is determined by (4.247) with the lower
(+) sign].
A Brane 2 model at low redshifts also masquerades as LCDM but with a larger value of
the density parameter. In this case,
LCDM
m
>
m
with
LCDM
m
being determined by
(4.247) with the upper () sign.
The range of redshifts over which this cosmic mimicry occurs is given by 0 z z
m
,
with z
m
determined by (4.244). Specically,
z
m
=
__
1 +

_
2/3

1/3
m
1 , (4.249)
which can also be written as
(1 +z
m
)
3
=

m
(1 +

b
)
(
LCDM
m
)
2
(4.250)
for both braneworld models.
In view of relation (4.250), it is interesting to note that we can use the equations de-
rived above to relate the three free parameters in the braneworld model: {

b
,
m
} to
164 4 The Universe with Matter and Dark Energy
Figure 4.35: An illustration of cosmic mimicry for the Brane 2 model. The Hubble parameter in three
low-density Brane 2 models with
m
= 0.04 is shown. Also shown is the Hubble parameter in a LCDM
model (dotted line) which mimics this braneworld but has a higher mass density
LCDM
m
= 0.2 (

=
0.8). The brane matter-density parameter (
m
) and the corresponding parameter of the masquerading
LCDM model are related as
m
=
LCDM
m

_
1
_

/ (1 +

b
)
_
, so that
m

LCDM
m
. The
redshift interval during which the braneworld masquerades as LCDM, so that H
BRANE2
= H
LCDM
for
z z
m
, is z
m
= 8, 32, 130 (left to right) for the three braneworld models.
_

m
, z
m
,
LCDM
m
_
. These relations (which turn out to be the same for Brane 1 and Brane 2
models) are
1 +

LCDM
m
=

LCDM
m

m
(1 +z
m
)
3
, (4.251)

LCDM
m
=
_
_

LCDM
m

LCDM
m
_
_
2
(1 +z
m
)
3
. (4.252)
Furthermore, if we assume that the value of
LCDM
m
is known (say, from the analysis of SN
data), then the two braneworld parameters

and

b
can be related to the two parameters

m
and z
m
, so it might be more convenient to analyze the model in terms of
m
and z
m
(instead of

and

b
). We also note that, under condition (4.244), the brane tension ,
determined by (4.241), is positive for Brane 1 model, and negative for Brane 2 model.
Since the Hubble parameter in braneworld models departs from that in LCDM at interme-
diate redshifts (z > z
m
), this could leave behind an important cosmological signature espe-
cially since several key cosmological observables depend upon the Hubble parameter either
differentially or integrally. Examples include:
4.9 On the Origin of the Dark Energy 165
Figure 4.36: The age of the universe in the Brane 2 model is shown with respect to the LCDM value.
The mimicry redshift is z
m
= 4 so that H
brane
(z) H
LCDM
(z) at z 4. The braneworld models
have
m
= 0.2, 0.1, 0.04 (bottom to top) whereas
LCDM
m
= 0.3. Note that the braneworld models are
considerably older than LCDM.
the luminosity distance d
L
(z):
d
L
(z)
1 +z
= c
_
z
0
dz

H(z

)
, (4.253)
the angular-size distance
d
A
(z) =
c
1 +z
_
z
0
dz

H(z

)
, (4.254)
the product d
2
A
(z)H
1
(z), which is used in the volume-redshift test
the deceleration parameter:
q(z) =
H

(z)
H(z)
(1 +z) 1 , (4.255)
the effective equation of state of dark energy:
w(z) =
2q(z) 1
3 [1
m
(z)]
,
m
(z) =
m
_
H
0
H(z)
_
2
(1 +z)
3
, (4.256)
the age of the universe:
t(z) =
_

z
dz

(1 +z

)H(z

)
, (4.257)
166 4 The Universe with Matter and Dark Energy
the electron-scattering optical depth to a redshift z
reion
(z
reion
) = c
_
z
reion
0
n
e
(z)
T
dz
(1 +z)H(z)
, (4.258)
where n
e
is the electron density and
T
is the Thomson cross-section describing scatter-
ing between electrons and CMB photons.
4.9.4 Effects from Inhomogeneous Universe
Recently, there have been several papers discussing the possibility that the apparent accel-
erated expansion of the universe is not caused by this mysterious dark energy, but rather by
inhomogeneities in the distribution of matter [10, 26]. The basic idea behind this line of ex-
planation is that we live in an underdense region of the universe, and the evolution of this
underdensity is what we percieve as an accelerated expansion. An analysis of early supernova
data by Zehavi et al. gave the rst indications that there might indeed exist such an underdense
bubble centered near us.
Specic models that give rise to such underdensities have been studied previously in the
form of a local homogeneous void. In these works both the underdensity and the region out-
side it are assumed to be perfectly homogeneous Friedmann-Robertson-Walker (FRW) models
with a singular mass shell separating the two regions. The inhomogeneity manifests itself as
a discontinuous jump at the location of the mass shell.
Figure 4.37: Backreaction in the inhomogeneous Universe.
In a homogeneous universe, it is possible to infer the time evolution of the cosmic expan-
sion from observations along the past light cone, since the expansion rate is a function of time
only. In the inhomogeneous case, however, the expansion rate varies both with time and space.
4.10 Summary 167
Therefore, if the expansion rates inferred from observations of supernovae are larger for low
redshifts than higher redshifts, this must be attributed to cosmic acceleration in a homoge-
neous universe, whereas in our case it can simply be the consequence of a spatial variation,
with the expansion rate being larger closer to us.
However, the supernova observations are not the only data that support the claim of an
accelerating expansion. As mentioned above, CMB observations also seem to lend support to
this claim. Therefore, in order for a model to be considered realistic, it should also be able to
explain the observed CMB temperature power spectrum. It is possible to obtain a very good
match to both the supernova data and the location of the rst acoustic peak simultaneously. In
fact, the match to the supernova data is better than for the CDM model.
The observed isotropy of the CMB radiation implies that we must be located close to the
center of the inhomogeneity. According to this picture, we are positioned at a rather special
place in the Universe. On the other hand, this model has the attractive feature that there is
no need for dark energy. Also the model is sufciently simple that it can be solved exactly.
It is therefore a good toy model for testing the ideas of inhomogeneities as a solution to the
mystery of the dark energy.
4.10 Summary
The time evolution of a Friedman Universe follows from the Friedman equations for
H =

R/R
H
2
=
8G
3

kc
2
R
2
+
c
2

3
(4.259)
and

R
R
=
4G(c
2
+ 3P)
3
+
c
2

3
. (4.260)
The present state of the FRW Universe is determined by 5 parameters:
1. H
0
= (

R/R)
0
: the Hubble constant
2.
m
=
m
/
crit
: nonrelativistic matter density
3.
k
= kc
2
/R
2
0
: curvature parameter
4.

=
DE
/
crit
: Dark Energy parameter
5. w
0
= P
DE
/
DE
1: EoS for Dark Energy.
Only two of the density parameters are independent

m
+
k
+

= 1 . (4.261)
The present concordance model has only two free parameters H
0
and
m
, since
k
= 0
and

= 1
m
,
a(t) =
_
_

m
1
m
sinh
_
3

1
m
H
0
t
2
_
_
2/3
(4.262)
168 4 The Universe with Matter and Dark Energy
Models can be tested with the following relations following from the Friedman equation
H = H(z)
1. the luminosity distance d
L
(z):
d
L
(z)
1 +z
= c
_
z
0
dz

H(z

)
, (4.263)
2. the angular-size distance
d
A
(z) =
c
1 +z
_
z
0
dz

H(z

)
, (4.264)
3. the product d
2
A
(z)/H(z), which is used in the volume-redshift test
4. the deceleration parameter:
q(z) =
H

(z)
H(z)
(1 +z) 1 , (4.265)
5. the effective equation of state of dark energy:
w(z) =
2q(z) 1
3 [1
m
(z)]
,
m
(z) =
m
_
H
0
H(z)
_
2
(1 +z)
3
, (4.266)
6. the age of the universe:
t(z) =
_

z
dz

(1 +z

)H(z

)
, (4.267)
7. the electron-scattering optical depth to a redshift z
reion
(z
reion
) = c
_
z
reion
0
n
e
(z)
T
dz
(1 +z)H(z)
, (4.268)
where n
e
is the electron density and
T
is the Thomson cross-section describing
scattering between electrons and CMB photons.
Bibliography
[1] P. Astier, J. Guy, N. Regnault et al.: 2005, The Supernova Legacy Survey: Mea-
surement of
M
,

and w from the First Year Data Set, astroph/0510447


[2] Branch, D., Tammann, G.A.: 1992, Ann. Rev. Astron. Astrophys. 30, 359
[3] Caldwell et al.: 2003, Physical Review Letters, 15 August
[4] Deffayet C.: 2001, Phys. Lett. B 502 199 [hep-th/0010186]
[5] A.E.Evrard, T.MacFarland, H.M.P.Couchman, J.M.Colberg, N.Yoshida, S.D.M.
White, A.Jenkins, C.S.Frenk, F.R.Pearce, G. Efstathiou, J.A.Peacock, and
P.A.Thomas: 2002, Galaxy clusters in Hubble Volume Simulations, ApJ 573, 7
[6] Friedmann, A.: 1922, Z. f.Physik 10, 377
Bibliography 169
[7] Hamuy, M. et al.: 1996, Astron. J. 112, 2408 (CTSS)
[8] Hamuy, M. et al.: 2001, Astrophys. J. 558, 615
[9] Knop, R.A. et al.: 2003, New Constraints on
M
,

, and w from an Independent


Set of Eleven High-Redshift Supernovae Observed with HST, ApJ 598, 102
[10] Kolb, E.W. et al.: 2005, hepth/0503117
[11] Mattig, W.: 1958, Astron. Nachr. 284, 109
[12] Maartens R.: 2004, Living Reviews Relativity 7 7 [gr-qc/0312059]
[13] Nesseris, S., Perivolaropoulos, L.: 2005, Comparison of the Legacy and Gold SnIa
Dataset Constraints on Dark Energy Models, astroph/0511040
[14] Perlmutter, S. et al.: 1999, Astrophys. J. 517, 565 (SCP)
[15] Perlmutter, S.A.: 2003, Physics Today
[16] Perlmutter, S., Schmidt, Brian P.: 2003, Measuring Cosmology with Supernovae,
astroph/0303428
[17] Riess, A.G. et al.: 1999, Astron. J. 117, 707 (HZSNS)
[18] Riess, A.G. et al.: 2004, Gold Sample: Type Ia Supernova Discoveries at z1 From
the Hubble Space Telescope: Evidence for Past Deceleration and Constraints on
Dark Energy Evolution, ApJ 607, 665; astroph/0402512
[19] Sahni V. 2005, Cosmological surprises from Braneworld models of dark energy,
Preprint astro-ph/0502032
[20] Schmidt, B. et al.: 1998, Astrophys. J. 507, 46
[21] Shtanov Yu V, On brane-world cosmology, 2000 Preprint hep-th/0005193
[22] Springel, V., White, S.D.M. et al.: 2005, Simulations of the formation, evolution
and clustering of galaxies and quasars, Nature, 2 June 2005
[23] Tonry, J.L. et al.: 2003, Cosmological Results from High-z Supernovae, ApJ 594, 1
[24] Vishwakarma, R.G.: 2005, Recent Supernovae Ia observations tend to rule out all
the cosmologies, astroph/0511628
[25] Weller, Albrecht, : 2002, ApJ
[26] Wiltshire, D.L.: 2005, grqc/0503099
List of Figures
1.1 The Aristotelian Universe . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.1 Universe scales . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.2 Superclusters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.3 Cepheid luminosity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.4 Hubble expansion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.5 Hubble expansion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.6 SNIa Hubble diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.7 SNAP Project . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.8 SNIa Hubble diagram from SNLegacy program . . . . . . . . . . . . . . . . 21
2.9 CMB Planck distribution . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.10 COBE dipole anisotropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.11 WMAP temperature ripples . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.12 Spherical harmonics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.13 Power spectrum of temperature uctuations before WMAP . . . . . . . . . . 25
2.14 WMAP power spectrum . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
2.15 DASI polarisation maps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.16 Triangulation of spheres . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.17 Galaxy distribution in CfA survey . . . . . . . . . . . . . . . . . . . . . . . 30
2.18 Galaxy distribution in 2dF galaxy survey . . . . . . . . . . . . . . . . . . . . 31
2.19 Redshift distribution in the 2dF survey . . . . . . . . . . . . . . . . . . . . . 31
2.20 SDSS vs CfA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
2.21 Abell 2104 at redshift 0.15 . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
2.22 Hydra cluster . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
2.23 Abell 2218 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
3.1 Einsteins Theorie der Relativitat von 1916 . . . . . . . . . . . . . . . . . . 48
3.2 Light cones . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
3.3 Light cone structure near a Black Hole . . . . . . . . . . . . . . . . . . . . . 63
3.4 Periheldrehung im Zweikorperproblem . . . . . . . . . . . . . . . . . . . . 65
3.5 Gravitational lense . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
3.6 Photon trajectories around a Black Hole . . . . . . . . . . . . . . . . . . . . 66
3.7 Slicing of SpaceTime . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
3.8 Spaces of constant curvature . . . . . . . . . . . . . . . . . . . . . . . . . . 72
3.9 World lines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
3.10 Observations in FRW . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
3.11 Distance ladder for the Hubble constant . . . . . . . . . . . . . . . . . . . . 80
172 List of Figures
3.12 Light bundle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
4.1 Classical Friedman models . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
4.2 Accelerated Universe . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
4.3 Expansion of the Universe with dark energy . . . . . . . . . . . . . . . . . . 106
4.4 Flat solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
4.5 Age of the Universe as a function of redshift . . . . . . . . . . . . . . . . . . 108
4.6 Age of the Universe . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
4.7 Causality problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
4.8 Equatorial slice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
4.9 Zoom in of the region . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118
4.10 Galaxies and quasars in the equatorial slice, displayed in lookback time coor-
dinates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
4.11 Zoom in of the region marked . . . . . . . . . . . . . . . . . . . . . . . . . 121
4.12 Square comoving grid shown in lookback time coordinates . . . . . . . . . . 122
4.13 Luminosity distance as a function of redshift . . . . . . . . . . . . . . . . . . 123
4.14 Luminosity distance as a function of redshift . . . . . . . . . . . . . . . . . . 125
4.15 Reduced distance modulus . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
4.16 Hubble diagram for SN Ia . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128
4.17 Hubble diagram for the brightest ellipticals . . . . . . . . . . . . . . . . . . 130
4.18 Light curve stretching of SNIa . . . . . . . . . . . . . . . . . . . . . . . . . 132
4.19 Distance modulus for SNIa . . . . . . . . . . . . . . . . . . . . . . . . . . . 133
4.20 Distance modulus for SNIa . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
4.21 Hubble diagram for cosmological Supernovae . . . . . . . . . . . . . . . . . 138
4.22 Fundamental plane for cosmological Supernovae . . . . . . . . . . . . . . . 139
4.23 Fundamental plane for cosmological Supernovae . . . . . . . . . . . . . . . 140
4.24 The difference in distance modulus between the empty model and the super-
nova and the GRB binned data . . . . . . . . . . . . . . . . . . . . . . . . . 145
4.25 Expected SNAP accuracy . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145
4.26 Future constraints on SN . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146
4.27 Apparent angular width . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148
4.28 Apparent angular width . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
4.29 Clustering of Dark Matter . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150
4.30 Comoving cosmic volume . . . . . . . . . . . . . . . . . . . . . . . . . . . 151
4.31 Number counts in the expanding Universe . . . . . . . . . . . . . . . . . . . 153
4.32 Empirical Fundamental Plane of Cosmology . . . . . . . . . . . . . . . . . . 155
4.33 Matter densities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158
4.34 Bulk and Brane in Braneworld . . . . . . . . . . . . . . . . . . . . . . . . . 160
4.35 Cosmic mimicry for the Brane2 models . . . . . . . . . . . . . . . . . . . . 164
4.36 Braneworld age . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165
4.37 Inhomogeneous Universe . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
List of Tables
3.1 Number of linearly independent pforms for D = 3 and D = 4. . . . . . . . 56
3.2 Distance modulus Virgo cluster . . . . . . . . . . . . . . . . . . . . . . . . . 79
4.1 Co-moving radii for different redshifts . . . . . . . . . . . . . . . . . . . . . 115
4.2 Fits of different cosmologies to available data sets . . . . . . . . . . . . . . . 142

You might also like