Sie sind auf Seite 1von 24

Cosmology Primer

Sean Carroll
http://preposterousuniverse.com/writings/cosmologyprimer/index.html

Cosmology Primer
Twentieth-century science completely overturned our view of cosmology. We now know that our solar system is one of
many in our galaxy, and our galaxy is one of many in the universe. These galaxies are spread throughout space in a nearly
uniform distribution, and distant galaxies are mutually moving apart from each other as the universe expands. Over ten
billion years ago the entire collection emerged from an incredibly hot and dense state: the Big Bang.
In the twenty-first century, new discoveries are offering new challenges to our understanding. Ordinary matter comprises
only about four percent of the stuff in the universe; the rest is dark matter and dark energy. Structures in the universe grew
from a very smooth primeval state; the tiny deviations from perfect smoothness may have been caused by a period of
inflationary expansion in the very early universe. New experiments are being designed to extend our understanding further
into the unknown.
This primer provides a brief introduction to these ideas, the basic picture of modern cosmology. The intended audience
includes anyone with curiosity about science; no technical background is assumed.
Contents:

Overview

A brief summary of the major features of our universe.


The Expanding Universe
The basic picture of modern cosmology: a smooth, expanding spacetime.
The Evolving Universe
A schematic timeline of major events, from the Big Bang to today.
The Luminous Universe
What we see when we look into the sky, from galaxies to cosmic rays.
The Dark Universe
Most of the stuff in the universe is unseen: dark matter and dark energy.
The Early Universe
Remnants of the Big Bang: the microwave background and primordial nuclei.
The Really Early Universe
Inflation, baryogenesis, and the beginning of time.
The Measured Universe
The instruments and techniques we use to do observations and experiments.
FAQ
Frequently Asked Questions about cosmology.

Cosmology Primer
I- Overview
The universe is a big place, filled with surprises, and including realms far beyond everyday experience. To be honest,
it's a lot to contemplate in one sitting. On this page of the Primer, we give a quick overview of the basic features of
our universe, as understood by contemporary cosmologists. The pages to follow go into more detail about the specific
ideas mentioned here.

The Milky Way galaxy, our home.


(Click image to enlarge.)

The most important features of the universe on very large scales can be summarized quite concisely: it's big, it's
smooth, and it's getting bigger. When we look into the sky on a clear night, the first things we notice (besides the
Moon and possibly the planets) are the stars. Each star is an object comparable to our Sun, although there is quite a
range of star types. We live in a collection of stars bound together by their mutual gravitational attraction: the Milky
Way galaxy.
But the Milky Way is just one of billions of galaxies in our observable universe; the basic characteristics of this
collection are described in the page on theluminous universe. Fortunately, on the largest scales the distribution of
galaxies is very similar throughout the universe; this is what is meant by the smoothness of the universe. Most
remarkably, distant galaxies are all moving away from each other, as discovered by Edwin Hubble in 1929. This
mutual moving-away is described in the page on the expanding universe.
The universe was a different place early in its history. Since the universe is expanding today, in the past it was much
more hot and dense, as its constituents were packed more tightly. Traced back sufficiently far, it should be too hot for
atoms to exist, as electrons are continually jostled away from atomic nuclei; even earlier, nuclei themselves would be
torn into individual protons and neutrons, and the universe should be an incredibly hot plasma of subnuclear particles.
If we try to go all the way to the beginning, we would reach a singular point called the Big Bang. A timeline of
important events in the history of the universe is provided in the page on theevolving universe.
Winding the picture forward from the Big Bang, we evolve from a hot and dense plasma to the cooled-off universe we

Cosmology Primer
observe today. Along the way, there is a brief period when the universe is about one minute old in which protons and
neutrons are being synthesized into light nuclei (helium, deuterium, and lithium, along with individual protons
comprising hydrogen nuclei). Later, electrons combine with these nuclei into atoms when the universe is about
370,000 years old. It is at this point that the universe becomes transparent, and the ambient radiation streams freely to
us today. Light waves get stretched to longer wavelengths as the universe expands, so this radiation has cooled to
approximately 3 degrees Kelvin, forming the cosmic microwave background (CMB). This basic framework is verified
to high precision by observations today, as described on the page about the early universe.

The inferred dark matter distribution (in blue) in a cluster of galaxies.


(Click image to enlarge.)

There is much more to the universe than what is immediately visible. Fortunately, everything gives rise to a
gravitational field, so we can detect additional sources of matter and energy by probing for their gravitational effects.
It turns out that only about 5% of the universe is "ordinary matter" (the atoms of which stars, gas, dust, and planets are
made). About 25% of the universe is "dark matter," a new kind of particle never yet detected in any laboratory here on
Earth. A full 70% of the universe is "dark energy," a smoothly-distributed and persistent kind of energy which remains
a fundamental mystery of science. The dark energy might be a vacuum energy inherent in spacetime itself, or it might
be something even more exotic. The pieces of evidence we have accumulated for dark matter and dark energy, as well
as theories proposed to account for them and experiments ongoing to constrain them, are discussed in the page on
the dark universe.
The conventional Big Bang model provides an excellent fit to observations, as long as we impose very specific initial
conditions at early times: an expanding universe with almost the same density at all points in space, but with small
perturbations that eventually grow into galaxies today. Why did the universe start like that? A possible answer is
provided by "inflation," which posits a brief period of accelerated expansion at very early times. A related problem
concerns the imbalance of matter and antimatter in the universe: why are there more particles than antiparticles? The
page on the really early universe discusses the idea of inflation as well as mechanisms to explain the matter/antimatter
asymmetry.
It goes without saying that we still have a lot to learn about the universe, but scientists are planning an exciting array
of new experiments to push the frontiers of our understanding forward. The page on the measured universe describes
the different kinds of experiments under development, as well as some of the most important ones already in
operation. We have every reason to believe that new discoveries about the universe will continue to surprise and
delight us

Cosmology Primer
II- The Expanding Universe
In thinking about the expanding universe, it is tempting to appeal to some sort of simile: distant galaxies are like
raisins in a baking loaf of bread, or dots drawn on the surface of a balloon. But the universe is a unique place, and
similes tend not to do it justice (or worse, to suggest something misleading). It's actually best just to think about the
universe itself, and what it looks like.
So imagine standing outside on a clear night and looking into the sky. Imagine further that you have perfect vision,
including not only ordinary light but all other kinds of radiation (radio waves, infrared and ultraviolet light, X-rays
and gamma-rays). The first thing you notice are stars; each star is much like our Sun, but further away and
correspondingly fainter. But the stars aren't distributed equally throughout the sky; they are arranged into a disk, and
our solar system is near one edge of the disk. This disk of stars, orbiting slowly under their mutual gravitational
attraction, is our galaxy, the Milky Way. There are almost one trillion stars in the Milky Way; in the night sky it shows
up as a faint band stretching from one horizon to the other.
But stars aren't the only thing we see; there are tiny patches of fuzzy light, which stand out in contrast to the pointlike
stars. The patches are "nebulae", and were a source of controversy earlier in the twentieth century -- were they clouds
of gas and dust within our galaxy, or separate galaxies in their own right? Eventually Edwin Hubble showed that many
(although not all) of the nebulae were in fact distant systems of stars, comparable in size to our own Milky Way
galaxy. In our observable universe, there are approximately one hundred billion such galaxies. For more discussion,
see the page on the luminous universe.

Galaxy map from the Sloan Digital Sky Survey.


(Click image to enlarge.)

Cosmology, as the study of the entire universe, would be a completely intractable subject if it weren't for one crucially
simplifying feature: if we look over large enough distances, the distribution of galaxies is basically the same
everywhere. In technical terms, we say that the universe is both "homogeneous" (the same at every point) and
"isotropic" the same in every direction). Of course these statements are not strictly true; the center of a galaxy has a
higher average density than the space in between galaxies. But as we look over larger and larger distances, the
deviations from place to place become smaller and smaller; once we are considering distances of hundreds of millions
of light years and more, the universe looks extremely uniform.
However, although galaxies are spread evenly throughout space, they are not static as a function of time; Hubble's
second great discovery is that the universe is expanding. The concept of an expanding universe can be a tricky one, so
it is worth being careful about what we mean. It is best to think of space itself stretching, so that the amount of space
between any two distant galaxies is increasing. The observable phenomenon that leads us to this conclusion is the
redshift: as light travels from one galaxy to another, its wavelength is stretched as the universe expands, reaching the
second galaxy having been shifted to the red (longer wavelengths). We therefore see relatively nearby galaxies slightly
redshifted, and very distant galaxies extremely redshifted. The cosmological redshift is similar in result (although
different in underlying cause) to the well-known Doppler shift that results when an object is moving away from you. It

Cosmology Primer
is therefore convenient to assign a "velocity" to this redshift; Hubble's Law states that this apparent velocity is directly
proportional to the distance to the galaxy, with the constant of proportionality being the Hubble constant. (At a deep
level, the expansion of space is different than the motion of objects through a static space; however, the differences are
negligible when the apparent velocities are much less than the speed of light, so the abuse of language is acceptable.)
Since we see distant galaxies moving away from us, it is tempting to think that we are in the center of something big.
But that's an incorrect impression; if we were living on any one of the other galaxies, we would still see all the
galaxies moving directly away from us, as a consequence of the general expansion. There is no center to the universe,
nor any preferred location; it's basically the same everywhere. Likewise, the universe is not (so far as we know)
expanding "into" anything; it's just that the amount of space in our single universe is growing with time.
Einstein's general theory of relativity, which states that spacetime is curved and that curvature is what we perceive as
"gravity," provides a dynamical framework for understanding the expansion of the universe. In cosmology, the
curvature of spacetime comes from two contributions: the curvature of space by itself, and the expansion rate of the
universe. The curvature of space is the same throughout space, and can be positive, negative, or zero; recent
observations of the Cosmic Microwave Background indicate that it is close to zero. General relativity then relates the
expansion rate and spatial curvature to the energy density of the universe -- the amount of energy in each volume of
space. If we know the spatial curvature, and we know the energy density, and we know how the energy density
changes as the universe expands, we can reconstruct the entire expansion history of the universe.
In particular, we can extrapolate backwards from our current situation to describe the very early universe. Since it is
expanding now, it was smaller in the past; galaxies were closer together, and the universe was both hotter and more
dense. Going back sufficiently far, the universe was so hot that galaxies and stars could not exist; further back,
individual atoms could not exist; further still, it was too hot for atomic nuclei themselves to exist. If we extend
ourselves fearlessly all the way back, the universe was of essentially zero size about 13.7 billion years ago -- the Big
Bang. Of course there is a lot we don't know about this period, although there is some that we do; see the pages on
the early universe and the really early universe for details.
If we can extrapolate into the past, we can also extrapolate into the future. The problem there, of course, is that we
have no observational data to check our speculations. Strictly speaking, therefore, we really don't know what the far
future history of the universe will bring. Our best current models, in which the universe is dominated by dark matter
and dark energy, predict that it is likely for the universe to continue expanding forever, becoming increasingly cold
and dark as time goes by. See the page on the dark universe for more specifics. However, our current ideas are still
speculative, so it pays to keep an open mind.

Cosmology Primer
III- The Evolving Universe
time = 10 -43 sec

size = 10 -30 today

temp = 10 32 Kelvin

The Planck era. Quantum gravity is important; current theories are inadequate. We can't get any closer to the Big
Bang at t=0 and say anything with confidence (or even with informed speculation).

time = 10 -35 sec

size = 10 -26 today

temp = 10 28 Kelvin

Inflation. A temporary period of domination by a form of dark energy at an ultra-high energy scale. A speculative
theory, but one that has so far been consistent with observations.

time = 10 -12 sec

size = 10 -15 today

temp = 10 15 Kelvin

Electroweak phase transition. At high temperatures, electromagnetism is unified with the weak interactions. This is
the temperature at which they become distinct.

time = 10 -6 sec

size = 10 -12 today

temp = 10 12 Kelvin

Quark-gluon phase transition. Quarks and gluons become bound into the protons and neutrons we see today.

time = 10 sec

size = 10 -9 today

temp = 10 9 Kelvin

Primordial nucleosynthesis. The universe cools to a point where protons and neutrons can combine to form light
atomic nuclei, primarily Helium, Deuterium, and Lithium.

time = 3.710 5 years

size = 10 -3 today

temp = 310 3 Kelvin

Recombination. The universe cools to a point where electrons can combine with nuclei to form atoms, and becomes
transparent. Radiation in the Cosmic Microwave Background is a snapshot of this era.

time = 10 8 years

size = 10 -1 today

temp = 30 Kelvin

The dark ages. Small ripples in the density of matter gradually assemble into stars and galaxies.

Cosmology Primer
time = 910 9 years

size = 510 -1 today

temp = 6 Kelvin

Sun and Earth form. From the existence of heavy elements in the Solar System, we know that the Sun is a secondgeneration star, formed about five billion years ago.

time = 13.710 9 years

size = 10 0 today

temp = 2.74 Kelvin

Today.

Cosmology Primer

IV- The Luminous Universe


Our Sun is a star, much like the other stars in the night sky except much closer to us. Astronomical distances are so
immense that scientists measure them in terms of the time it takes light to travel between objects. Light from the Sun
takes about eight minutes to reach the Earth, so we say the Sun is eight light-minutes away; this is equal to ninetythree million miles. The nearest star, in contrast, is over four light-years away -- almost twenty-four trillion (2410 12)
miles.
Stars are not distributed evenly throughout the visible universe. They are grouped into galaxies, collections of millions
of stars orbiting each other under their mutual gravitational attraction. Our galaxy, the Milky Way, has approximately
one hundred billion (1011) stars. Many of these stars are likely to have their own planets orbiting around them;
astronomers have discovered numerous planets outside the Solar System, although not a sufficient number to draw
reliable conclusions about their abundance. In addition to the stars and planets, galaxies contain large clouds of gas
and dust, often in the process of collapsing to form new stars.

Galaxies in the Hubble Deep Field.


(Click image to enlarge.)

Galaxies come in many different forms, depending on their age and history as well as the distribution of stars, gas, and
dust. The image on the right is part of the Deep Field image from the Hubble Space Telescope, and shows a wide
variety of galaxies. Among the different kinds of galaxies, three varieties are especially fundamental: spirals,
ellipticals, and irregulars. Spiral galaxies are large disks containing both stars and dust clouds, typically orbiting
around a central bulge of stars. Seen face-on, these galaxies exhibit dramatic spiral arms, indicating regions where
stars are being formed. Elliptical galaxies, in contrast, are dense with stars but have relatively little dust, and are
similar to the central bulges of the spirals. Irregular galaxies do not feature the well-defined shapes of spirals and
ellipticals, and are often smaller galaxies.
While stars orbit each other in galaxies, galaxies themselves often orbit within collections of other galaxies. Our own
Milky Way has a number of satellite galaxies, including the Magellanic Clouds visible in the Southern Hemisphere
sky. There are also larger collections of galaxies: groups, clusters, and superclusters. The Milky Way is a member of
the Local Group, in which the other large member is the Andromeda Galaxy, M31. Superclusters are the largest
gravitationally bound systems, but the distribution of galaxies on scales larger than superclusters is nevertheless not
completely uniform; the deviations from perfect regularity form the large-scale structure of the universe. The galaxy
map from the Sloan Digital Sky Survey shows evidence of large-scale structure.

Cosmology Primer
Stars and galaxies are the most obvious features of the universe, at least if we are observing it using ordinary visible
light. In an effort to learn as much as we can, astronomers will often turn to other forms of light. Light can be thought
of as either an electromagnetic wave, or as individual particles called "photons." Radiation with a longer wavelength
than visible light (infrared and radio waves) correspond to lower-energy photons, while shorter wavelengths
(ultraviolet light, X-rays and gamma rays) are high-energy photons.
Observations in different wavelengths give evidence of a violent universe in a constant state of flux. Infrared and
radio observations show regions of gas and dust collapsing to form new stars. X-rays and gamma rays, at shorter
wavelengths and thus higher frequencies, are formed by very high-energy processes in extreme environments. It is
from such observations that we infer the existence of black holes -- regions of spacetime where the gravitational field
is so strong that nothing entering can ever escape. In the observed universe, black holes seem to come in two major
types: supermassive black holes (over a million times the mass of the Sun) at the centers of large galaxies, and smaller
black holes (a few times the mass of the Sun) scattered throughout our galaxy. We suspect that these smaller black
holes result from supernovae -- the explosions of stars at the end of their life cycles. Supernovae are rare (perhaps one
per galaxy per century), but by observing thousands of galaxies at once we are guaranteed to discover a respectable
number on demand; surveys of just this sort have been crucial in making the case for dark energy, as explained in the
page on the dark universe.
Most of the photons in the universe are actually not emitted by stars, gas, or any other object in the contemporary
universe. Rather, they are the low-energy photons left over from the Big Bang, as described in the page on the early
universe. This relic radiation is primarily in the microwave region of the spectrum, and is known as the Cosmic
Microwave Background (CMB). Photons from the CMB pervade space, providing a background buzz that serves as a
constant reminder of the hot, dense state in which our universe began.
Since ancient times, almost everything we have learned about the universe has come through observing light of
various forms. Modern astronomers, however, are increasingly taking advantage of other ways to probe the universe.
In addition to ordinary photons, numerous other kinds of particles are constantly bombarding the Earth: protons (and
atomic nuclei) that make up cosmic rays, ultra-light neutral particles known as neutrinos, and the gravitational
equivalent of electromagnetic waves, known simply as gravitational waves. Cosmic rays have proven to be an
invaluable window onto energetic processes in the universe, leading to important insights into particle physics (such
as the discovery of the muon, a heavier cousin of the electron); today, the origin of the highest-energy cosmic rays
remains a deep mystery. Neutrinos from astrophysical sources have likewise led to significant discoveries; it was the
shortage of neutrinos emitted from the Sun which gave the first clue that these particles might have small masses,
rather than being strictly massless. Gravitational waves have never been detected directly, although new observatories
are coming online with the goal of doing exactly that; however, the effect of gravitational waves has been observed in
the loss of energy from orbiting neutron stars, for which Hulse and Taylor were awarded the Nobel Prize.

Cosmology Primer
V- The Dark Universe
Our knowledge of the universe comes from looking at it in various ways -- different wavelengths of light, neutrinos,
cosmic rays, and hopefully some day gravitational radiation. But how do we know that we are seeing everything there
is? How could we determine whether there were substances in the universe not directly visible in our telescopes?
The answer lies with gravity. Long ago, Galileo determined that every object falls the same way in a gravitational
field. Later, Einstein extended this idea: every substance with any kind of energy will create a gravitational field.
(Note that mass is a kind of energy, since E=mc2.) So we can, in principle, detect everything in the universe, by
mapping out the gravitational field throughout spacetime. To our surprise, there is much more out there than we can
see directly. The invisible stuff comes in two forms: dark matter and dark energy.
Dark matter is some kind of particle that we have not yet detected in experiments here on Earth, but nevertheless
comprises most of the matter in the universe. The first evidence for its existence came from studying the dynamics of
galaxies and clusters of galaxies. The basic point is that something in orbit around a massive object moves more
rapidly, the more mass the object has. Fritz Zwicky was the first to put this idea into action, studying the motion of
galaxies in the Coma cluster; their motions were too rapid to be accounted for by the visible matter in the galaxies.
Later, Vera Rubin looked at matter orbiting at the edges of individual galaxies, and noticed an similar effect -- the
rotation speeds of the galaxies did not fall off with distance as they should if the gravitational fields were being caused
by the visible matter alone.
These discrepant motions, and more modern methods confirming this behavior with greater precision, are strong
evidence for unseen matter in galaxies and clusters. Of course, it is natural to imagine that the extra matter is quite
ordinary, but simply invisible and transparent (much like air on a clear day). And indeed, observations in different
wavelengths have provided evidence for previously unseen gas in galaxies and clusters. However, we have very good
evidence that the substantial majority of dark matter is something exotic, rather than ordinary matter that we can't see.
This evidence comes from two independent sources -- the abundance of light elements from primordial
nucleosynthesis, and temperature anisotropies in the cosmic microwave background. Both phenomena are described
in the page on the early universe. The result is that the abundance of exotic dark matter, some kind of particle that has
never been directly observed here on Earth, is about five times the abundance of ordinary matter in the universe.
What we know about dark matter, then, is that it is a new kind of unseen particle that falls readily into galaxies and
clusters. The fact that it is dark means that it is neutral, rather than electrically charged; in general, charged particles
interact readily with light, and would not be dark. The fact that the dark matter is concentrated in galaxies and clusters
indicates that it is slowly moving; particles of this form are referred to as "cold." What could the cold dark matter be
made of? We are not at all sure, although numerous theories have been proposed. Two ideas are especially popular:
neutralinos and axions. "Neutralinos" are a kind of particle predicted by supersymmetry, a popular (but as yet purely
conjectural) theory in particle physics. According to supersymmetry, each kind of known particle has a "superpartner"
with a different intrinsic spin; the neutralino is simply a massive, neutral, stable superpartner of one of the known
particles, an is a natural dark matter candidate. (Such a particle would interact predominantly through the weak
nuclear force, and is therefore an example of a Weakly Interacting Massive Particle, or WIMP.) Axions, on the other
hand, are another kind of hypothetical particle, originally postulated to explain certain symmetries of the strong
nuclear force. In the case of both neutralinos and axions, active experimental programs are underway to detect these
dark matter candidates in the laboratory, either by producing them directly or by observing the effects of ambient
particles floating through the Solar System. Finally, it may be possible to detect dark matter particles indirectly, if they
annihilate into photons in high-density regions of the universe; the resulting radiation would have a characteristic form
that would signal the existence of a new kind of particle.
Of course, there is a kind of known particle which is neutral, stable, and (as we now know) massive -- the neutrino.
Neutrinos are abundant in the universe, approximately as abundant as photons. However, they are not good dark
matter candidates, because they are not very cold. Even if neutrinos today are moving relatively slowly, in the early
universe they were moving near the speed of light. As a result, they would stream freely away from galaxies and

10

Cosmology Primer
clusters, rather than settling into them as dark matter is observed to do. Indeed, a useful way to put an upper limit on
the mass of the neutrino is to insist that it not comprise such a significant fraction of the mass of the universe that it
would interfere with the evolution of large-scale structure.

A supernova at the edge of a distant galaxy.


(Click image to enlarge.)

As if dark matter weren't exotic enough, dark energy is even more mysterious. We know very little about dark energy,
other than two characteristic features: it is spread uniformly throughout space, and maintains an approximately
constant density as the universe expands. It must be (nearly) uniform throughout space, since otherwise it would
clump into galaxies and clusters and affect local motions just as dark matter does. Instead, the dark energy only affects
the overall curvature of spacetime. How, then, do we know that dark energy exists? Its affects on spacetime curvature
show up in two ways -- it makes the universe accelerate, and it contributes (along with matter) to the curvature of
space alone.
According to Einstein, the rate of expansion depends on the average energy density of the universe. To measure the
expansion rate, we use standard candles, objects whose intrinsic brightness is known. The further away a standard
candle is, the dimmer it will appear, allowing us to accurately determine its distance. We want standard candles that
are very bright, so they can be observed at cosmological distances; good candidates are provided by supernova
explosions, which can rival the entire light output of the galaxy they are in. It turns out that supernovae come in
various forms, of different brightness; but a certain kind, the Type Ia supernovae, have a brightness which depends
directly on how fast they explode and fade away. Observations of Type Ia supernovae at high redshifts by two groups
(the Supernova Cosmology Project and the High-Redshift Supernova Team) in 1998 provided the first direct evidence
that the universe is accelerating rather than slowing down.
Why does dark energy make the universe accelerate? Because, unlike matter and radiation, it does not dilute away as
the universe expands -- the density of dark energy remains close to constant. Therefore, according to Einstein, the
Hubble parameter remains close to constant. But remember that the apparent velocity of a galaxy is given by the
Hubble parameter times the distance, as explained in the page on the expanding universe. Since the distance to any
given galaxy is increasing, a nearly-constant Hubble parameter implies that the apparent velocity will also be
increasing -- in other words, the galaxies are accelerating away from us.
The other piece of evidence for dark energy comes from the curvature of space, as measured through temperature
fluctuations of the Cosmic Microwave Background. As described in the page on the expanding universe, the curvature
of spacetime can be thought of as a combination of the curvature of space, and the expansion of space through time.
Through observations of temperature fluctuations in the CMB (described in the page on the early universe), we can
measure the overall curvature of space, and find that it is close to zero. Meanwhile, the total amount of matter in the
universe (both ordinary and dark) falls well short of what is needed to explain the flatness of space. It turns out that

11

Cosmology Primer
the amount of dark energy required to explain the acceleration of the universe is just right to explain the fact that
space is flat. We therefore seem to have a complete inventory of the constituents of our contemporary universe:

five percent ordinary matter,


twenty-five percent dark matter,
seventy percent dark energy.
The completion of this inventory is one of the most impressive successes of modern cosmology.
We don't know what the dark energy is. Since observations constrain its density to be close to uniform throughout
space and approximately constant in time, the simplest candidate would be something that is exactly constant in space
and time. Such a substance is called vacuum energy -- an energy density that is inherent in empty space itself,
unchanging from point to point in the universe. But once we admit the possibility of vacuum energy, we can go back
and estimate how large such energy should be. The result, according to our best understanding of quantum field
theory, is 10120 times the observed amount -- a fantastically large factor, and a sure sign that there is something we
don't understand about vacuum energy. This discrepancy is known as the "cosmological constant problem," since the
vacuum energy turns out to be equivalent to Einstein's old idea of a cosmological constant. In addition to this problem,
there is the puzzle of why the dark energy and the matter density are relatively close to each other (only differing by a
factor of two or three. After all, matter dilutes away as the universe expands, while dark energy remains essentially
constant, so their relative abundance changes dramatically; we seem to be born lucky, in an era when both matter and
dark energy play an important role in the universe.
In an effort to address these puzzles, cosmologists are trying to learn all they can about dark energy. In particular, they
would like to know whether it really is absolutely constant, or merely changing very slowly. If dark energy is
dynamical, rather than a strictly-constant vacuum energy, its origin in fundamental physics would be completely
different. It is even conceivable that the acceleration and flatness of the universe are not due to dark energy at all, but
rather to a breakdown of Einstein's theory of general relativity on cosmological scales. All of these possibilities are in
play, and future observations will help us decide between them once and for all.

12

Cosmology Primer

VI- The Early Universe


Given our understanding of the current state of the universe, and our knowledge of the appropriate laws of physics, we
can extrapolate backwards in time to say what the early universe must have been like. Fortunately, we can then use
current observations to test whether such an extrapolation is valid; the answer is that it is remarkably accurate.

Our improving view of the cosmic microwave background, from 1992 to 2003. The top view is from the COBE satellite, and the bottom
from the more recent WMAP satellite.
(Click image to enlarge.)

The cosmic microwave background (CMB) is the leftover radiation from the Big Bang. When the universe was much
smaller it was much hotter and denser. The ordinary matter that today resides in the form of stars, gas, and dust was
packed together at incredible densities and temperatures, so much so that electrons moved freely rather than being
attached to individual atomic nuclei (much like conditions at the center of a star). This hot plasma gave off copious
amounts of radiation, just like any other hot object, and we can detect that radiation today. The plasma was also
opaque; any photon would only travel a short distance before bumping into a free electron. At 370,000 years after the
Big Bang, the temperature had dipped below about 3,500 Kelvin, cool enough for electrons to recombine with nuclei
to make atoms, and the universe suddenly became transparent. As the universe expanded and photons redshifted to
longer wavelengths, the radiation subsequently cooled to about 2.7 Kelvin, which is what we see today. The CMB was
first discovered by Arno Penzias and Robert Wilson in 1965.
The CMB provides a snapshot of what the universe looked like at the moment of recombination, when electrons
became attached to nuclei. What it looked like was something extremely smooth; fluctuations in density from place to
place were only about one part in 100,000. But we can detect these fluctuations. The image reproduced here is the
famous map from the Wilkinson Microwave Anisotropy Probe satellite, compared with the first detection of CMB
fluctuations from the earlier COBE satellite. Blue regions are slightly colder than average, red regions are slightly
hotter. These changes in temperature from place to place are referred to as "anisotropies," from the technical term
meaning "changing as we look in different directions." Of course the relic radiation from the Big Bang fills all points
of space, no matter where we are in the universe; but since we can only observe it from our location, we perceive the
the changes from place to place as projected onto the sphere of the sky.
The smooth, slightly perturbed early universe visible in the CMB anisotropies grew into the lumpier universe of stars
and galaxies we see today. This should come as no surprise. The hot and cold spots of the CMB correspond to regions
of slightly higher or lower density than average. In the regions that were overdense, the pull of gravity brought matter
closer together, further emptying out the regions that were less dense. The evolution of the universe under the

13

Cosmology Primer
influence of gravity thus acts to increase the contrast of the matter distribution, from a nearly featureless plasma to an
intricate collection of galaxies. This process takes longer over larger distances, which is why the universe remains
approximately smooth on very large scales.
There is a treasure trove of information contained in the CMB fluctuations. In particular, statistical properties of the
fluctuations depend on two things: the original primordial perturbations from which they arose, and the recipe of
ingredients in our universe that controls the subsequent evolution of the perturbations between early times and now.
Remarkably, an extremely simple specification of primordial perturbations works very well -- simply imagining that
the perturbations are (on average) of equal strength at all distance scales. From this guess, and the observed
fluctuations in the CMB sky, we can derive very tight constraints on interesting cosmological parameters. In
particular, the CMB provides independent support for the ideas that there is more matter in the universe than can be
accounted for by ordinary atoms (thus implying the need for dark matter), and that there is more total energy than be
accounted for by matter along (thus implying the need for dark energy).
The CMB provides a valuable picture of what the universe was like when it was 370,000 years old, but it also forms a
barrier past which we can't see -- because the plasma of the early universe was opaque, we can never obtain a direct
image of any event at earlier times. Nevertheless, we can obtain indirect information about the universe back to when
it was a few seconds old, by observing the primordial abundances of light elements.
Just as the universe at times earlier than 370,000 years was too hot for electrons and nuclei to stick together to form
atoms, at times earlier than a couple of minutes after the Big Bang it was so hot that protons and neutrons could not
stick together to form atomic nuclei. As it expanded and cooled, the early universe was a nuclear reactor -- protons
and neutrons combined to make light elements such as deuterium ("heavy hydrogen," with one proton and one neutron
per nucleus), helium (two protons and two neutrons, or occasionally two protons with only one neutron), and lithium
(three protons and four neutrons). The nuclei themselves would have liked to continue this process of fusion,
combining all the way to form iron (the most stable nucleus), but the rapid expansion of the universe soon made the
nuclear plasma too thin to sustain further reactions.
This process of primordial nucleosynthesis took place while the universe was between a few seconds and a few
minutes old, and only one-billionth its current size. Since that time, further nuclear evolution has taken place inside
stars and supernovae, providing us with the rich variety of elements in the universe today. However, by looking to
regions that have been relatively untouched by exploding stars, we can infer the primordial abundances of deuterium,
helium, and lithium. These abundances depend sensitively on two things -- the amount of ordinary matter in the form
of protons and neutrons, and the expansion rate of the universe when it was just a few seconds old. We obtain perfect
agreement with observations if two things are true -- the amount of ordinary matter is much less than the total amount
of matter we deduce in the current universe (thus providing evidence for dark matter), and the expansion rate is
exactly as predicted by Einstein's theory of general relativity (thus assuring us that our best theories can be safely
extrapolated back to early times).
The agreement between the predictions of the Big-Bang cosmological model and our observations of primordial lightelement abundances was by no means guaranteed -- it is certainly conceivable that the observed amounts of
deuterium, helium and lithium could not be explained by conventional general relativity for any value of the matter
density. The fact that we do get good agreement, indicating that we understand the behavior of the universe when it
was only a few seconds old, is one of the most profound achievements of modern science.

14

Cosmology Primer

VII- The Really Early Universe


We have no direct knowledge of what the universe was like before the Big-Bang nucleosynthesis era, when the
universe was between a few seconds and a few minutes old (as described in the page on the early universe). It is worth
emphasizing that any ideas we have about earlier times are only that -- ideas. Nevertheless, just as we can successfully
extrapolate the laws of physics from the present day back to the time of nucleosynthesis, we may also extrapolate
these laws even further back, to construct a picture of what the very early universe may have been like.
In a universe dominated by matter and radiation (as opposed to dark energy), the mutual gravitational pull of all the
particles tends to slow down the expansion rate as the universe expands. When the universe was smaller and more
dense, it therefore follows that the expansion rate was much larger than it is today. Indeed, as we extrapolate the
universe further back in time, we reach a point where the density, temperature, and expansion rate were all infinitely
large. This point is a singularity, which we refer to as the Big Bang (although that term is also used for the entire
cosmological model that includes the later universe as well). At the Big Bang, our knowledge of what happens gives
out; the fact that physical quantities become infinite is a sign that we don't know what is going on. Presumably, in the
real world there is no singularity; instead, something happens that cannot be described by physics as we currently
understand it.
Just because we don't understand the Big Bang itself doesn't mean we can't usefully talk about the period immediately
afterwards, when the universe was in a hot, dense, rapidly expanding state. In the absence of a sensible theory of the
origin of the universe, cosmologists ask what initial conditions are necessary to explain the observed features of our
universe today. But in fact we want more than that; we would like to believe that these initial conditions are somehow
natural, rather than arbitrarily finely-tuned. This desire may or may not be accommodated by reality, but has led to a
great deal of interesting speculation about the very early universe.
One puzzle we have about the universe is the apparent dominance of matter over antimatter. Every type of elementary
particle (electrons, protons, neutrons, and so on) has a corresponding type of antiparticle (positrons, antiprotons,
antineutrons) of equal mass and opposite electric charge. But what we observe in the universe is overwhelmingly
matter and not antimatter, which we know because matter and antimatter tend to explosively annihilate when they
come into contact with each other. If other galaxies, for example, were made of antimatter, there would be regions in
between where particles would intermix, giving rise to high-energy radiation that has not been detected. It is possible
that this asymmetry between matter and antimatter is simply a feature of the initial conditions of the universe, but it
would seem more satisfying if we could explain how it arose dynamically as the universe evolved. Such a
hypothetical process is known as "baryogenesis," since the observed imbalance between matter and antimatter is
actually an imbalance between baryons (protons and neutrons) and antibaryons. There are numerous models of
baryogenesis, many of which may be testable at upcoming particle accelerators; to date, however, no single model has
proven so successful that it has been accepted as a standard picture.
Another unusual feature of our universe is its smoothness -- the distribution of galaxies is uniform on large scales, and
the microwave background provides strong evidence that matter was even more smoothly distributed at earlier times.
This uniformity should strike us as unusual, since small deviations tend to grow with time, as overdense regions
collapse to form stars and galaxies. But the finite speed of light makes the situation even more surprising. The CMB
shows us what the universe was like 370,000 years after the Big Bang. But when we observe widely separated parts of
the CMB, we are seeing regions of the universe that were much more than 370,000 light-years apart at that time. In
other words, there was not enough time for any signal to travel from one region to another. So how do these separated
regions know that they should be at the same temperature? This conundrum is known as the "horizon problem," since
the finite distance light can travel since the Big Bang defines an horizon around each point, and the horizons of distant
parts of the microwave sky do not overlap.

15

Cosmology Primer
Unlike the puzzle of the baryon asymmetry, the horizon problem does have a popular solution -- the idea of inflation.
One way of thinking of inflation is to simply imagine that the very early universe went through a period where it was
temporarily dominated by an extremely large amount of dark energy, which then suddenly decayed into ordinary
matter and radiation. This inflationary dark energy caused the universe to accelerate at a fantastic rate, taking nearby
points and moving them very far apart -- so that the widely-separated regions we observe in the CMB were actually
quite nearby and in contact early on. This elegant solution to the horizon problem comes along with extra benefits. For
one thing, the process of inflation takes any initial curvature of space and diminishes it to near zero, explaining the
observed flatness of the universe. For another, inflation wipes out any particles that may have existed before inflation
began, which is useful if we don't observe those particles today. An example is provided by "magnetic monopoles,"
which some theories of particle physics predict should exist in copious amounts, even though none has ever been
detected. The desire to get rid of magnetic monopoles was actually the original motivation that led to the invention of
inflation by Alan Guth in 1980.
Inflation has another unanticipated benefit: it provides a possible origin of the primordial density perturbations that
lead to temperature anisotropies in the CMB and ultimately grow into the large-scale structure we observe today. This
phenomenon arises from considering inflation in the context of quantum mechanics. Modern physicists understand
that classical mechanics, in which particles have definite positions and velocities, is only an approximation to
quantum mechanics, in which such quantities are subject to a certain irreducible uncertainty. The same thing holds
true for the density of an expanding universe. Thus, while inflation does its best to make the universe absolutely
uniform, quantum mechanics prohibits it from doing so; there is always a small amount of fluctuation in the amount
of energy from place to place that no amount of inflation can erase. Indeed, we can use the rules of quantum
mechanics to predict what kinds of fluctuations should arise from inflation. The result is a set of perturbations of
approximately equal strength at all distance scales. As mentioned in the page on the early universe, these are precisely
the kind of fluctuations needed to explain what we observe in anisotropies of the CMB. This doesn't mean that
inflation is necessarily correct, but certainly provides some evidence in its favor.
But there remains a great deal that we don't understand about inflation. In particular, while the general framework
remains attractive, no specific model of inflation has become popular. In other words, we don't know exactly what this
mysterious dark energy was that dominated the universe at very early times, nor how it converted into ordinary matter
and radiation. A possible clue could come from another prediction of inflation: gravitational waves. Just as quantum
mechanics predicts irreducible fluctuations in the density of matter during inflation, it also predicts fluctuations in the
gravitational field, which manifest themselves as gravitational waves. These waves can lead to a specific unmistakable
signature in the polarization of the microwave background. Unfortunately, we don't know for sure how strong these
gravitational waves will be, and they might be so weak as to be undetectable. But cosmologists are planning
experiments to look for them, and if they are detected it will be a great triumph for inflation.
In a sense, inflation hides from view anything that came before it. Nevertheless, we are still curious about the very
origin of the universe, and the conditions that gave rise to inflation (if indeed it happened). Presumably any sensible
description of this epoch will involve quantum gravity (the long sought-after reconciliation of quantum mechanics
with Einstein's general relativity), and perhaps require an understanding of more esoteric physics such as superstring
theory.

16

Cosmology Primer

VIII- The Measured Universe


Astronomy, arguably the oldest science, has traditionally depended on observations of phenomena in the sky. The
observational nature of the discipline distinguishes it from experimental sciences like physics or chemistry, in which
we can control the conditions under which we measure the systems of interest, gradually altering configurations and
repeating the experiments as necessary. When it comes to phenomena outside our Solar System, we have to take what
the universe gives us. The difference between experimental and observational science is much like the difference
between asking someone questions versus eavesdropping on their conversations -- asking our own questions allows us
to be precise and ask for clarification, but eavesdropping can allow us to learn secrets that people would never have
revealed under direct interrogation.
In looking out at the universe, our most straightforward tool is ordinary light, the same messenger that allowed our
ancestors to chart the stars and planets. We now know, of course, that visible light is just one form of electromagnetic
radiation. Other forms stretch from very long wavelengths (infrared light and radio waves), through visible light,
down to very short wavelengths (ultraviolet light, X-rays, and gamma-rays). All of these wavelengths of light are
emitted by objects in the universe, and astrophysicists take advantage of all of them to get as complete a picture of the
universe as possible, as discussed in the page on the luminous universe.
Galileo, the first person to observe the sky through a telescope, demonstrated the existence of objects (such as the
moons of Jupiter) that would never have been found by the naked eye, and we continue to follow in his footsteps. The
basic principle of a telescope (working in ordinary visible or infrared light) is simply to collect a large number of
photons (light particles) and focus them to a detector. Each individual photon is distinguished by three characteristics:
its frequency (or equivalently wavelength, or energy), the direction from which it arrives, and the time at which it is
observed. Different detectors may keep track of some or all of this information; studying the distribution of photon
frequencies is spectroscopy, the precise directions from which the different photons arrive is of course imaging, while
studying the amount of light arriving as a function of time is photometry. Modern telescopes strive for better views of
the universe both by increasing the size of their light-collecting area -- typically with large mirrors -- and by being
located in places where interference from the Earth's atmosphere is minimized -- either in dry, high-altitude climates,
or somewhere in orbit outside the atmosphere entirely.
Telescopes using visible light continue to produce startling discoveries. The Sloan Digital Sky Survey (SDSS), for
example, makes use of a dedicated telescope at Apache Point in New Mexico to survey a large area of the deep sky for
galaxies and quasars as well as ordinary stars. The SDSS first images the sky, and then does spectroscopy on the most
interesting objects found so as to determine their cosmological redshift (as explained in the page on the expanding
universe). Since redshift is proportional to distance according to Hubble's Law, the result is a threedimensional map of the large-scale structure of the universe. Both in the SDSS and in future surveys, an important
goal will be to study large-scale images of galaxies to search for the slight distortions of images due to the
gravitational lensing of background objects by galaxies in the foreground.
In a complementary vein, a satellite such as the Hubble Space Telescope (HST) can peer extremely deeply into one
region of the sky to reveal galaxies in the early stages of their evolution. The great advantage of being in orbit is not
that you observe more photons, but rather that the absence of atmospheric distortion allows for the collection of more
precise and detailed images. Such improved resolution can be crucial, for example, in the study of distantsupernovae,
where it is important to distinguish the supernova event from the light of the surrounding galaxy. A successor to the
HST, the James Webb Space Telescope, is currently under development by NASA; it will be able to look even deeper
into the universe, to help us understand how early galaxies were assembled from relatively smooth distributions of gas
and dark matter.
At longer wavelengths, radio telescopes have become an indispensable part of observational astronomy. Several Nobel

17

Cosmology Primer
prizes have gone to discoveries made with radio telescopes, from the discovery of pulsars (rapidly rotating neutron
stars) to the first observations of the microwave background itself. Along with visible light, the radio band is one in
which radiation can readily penetrate the Earth's atmosphere, so ground-based telescopes are extremely valuable.
Nevertheless, just as with visible light, there is much to be gained by minimizing interference with the atmosphere. It
is therefore common to place radio telescopes in locations where the atmosphere is both thin and stable; popular
choices include the high plains of Chile, and the extreme cold of the South Pole. More dramatic methods of
minimizing atmospheric interference include placing radio telescopes on long-duration balloon flights, or simply
putting them on satellites completely above the atmosphere. These techniques are especially useful when observing
the microwave background; the CMB is at such a low temperature that tiny amounts of contamination from the Earth
itself can be troublesome to experiments. The Planck satellite, a planned collaboration between NASA and the
European Space Agency, will hopefully provide maps of the CMB of even higher precision than those produced to
date. Another goal is to obtain high-precision measurements of the polarization of the microwave background; as
explained in the page on the really early universe, such observations may provide crucial clues to the nature of
inflation.
In contrast to radio waves and visible light, higher-frequency waves of ultraviolet light and X-rays are unable to
penetrate through the atmosphere, and it is necessary to go into orbit to observe them directly. NASA's Chandra
satellite has provided an unprecedented view of the X-ray sky; future missions are planned that will inventory black
holes throughout the universe, map out hot gas in clusters of galaxies, and image the innermost regions close to black
holes themselves. Once we consider extremely high-energy gamma rays, however, the number of photons emitted by
interesting sources becomes very small. The planned GLAST satellite will feature a large area telescope to collect as
many gamma-ray photons as possible. Alternatively, we can take advantage of the fact that high-energy gamma rays
give off a kind of secondary (Cerenkov) radiation when they collide with the atmosphere. This radiation can be
detected, and from that we can reconstruct information about the original gamma rays; this technique has been
successfully used at the Whipple observatory in Arizona, and is the basis for its successor, the VERITAS observatory.
As mentioned in the page on the luminous universe, photons are not the only means by which we can observe the
universe. Other kinds of particles which we are able to observe include cosmic rays and neutrinos, with very different
techniques applicable in either case. Cosmic rays, which are thought to be individual protons or atomic nuclei that
have been accelerated to tremendous energies, are similar to very high-energy gamma rays, in that the number of
particles is extremely low. But because the energy of each particle is so large, it is again possible to observe secondary
effects when the cosmic rays interact with the atmosphere. The Pierre Auger observatory, which aims at understanding
cosmic rays at the very highest energies, uses two complementary techniques: direct detection of Cerenkov radiation
in an array of detectors on the ground, and observations of flashes in the atmosphere caused when secondary particles
cause nitrogen to fluoresce.
Neutrinos are also observed using several different techniques. One type of detector uses the fact that a neutrino can
interact with an atomic nucleus to turn it into a different element entirely, such as chlorine being converted to argon. It
was a detector of this type at the Homestake mine in South Dakota that first provided evidence for an anomalously
low flux of neutrinos from the Sun. Alternatively, a neutrino can knock an electron out of an atom, and the electron in
turn can travel through a medium such as water and give off detectable Cerenkov radiation. This technique is only
sensitive to higher-energy neutrinos, but has the advantage of providing information about the direction and timing of
the event; the Super-Kamiokande observatory in Japan used this technique to detect neutrinos from Supernova 1987a
in the Large Magellanic Cloud. An advanced version of this technique, used by the Sudbury Neutrino Observatory,
uses heavy water (in which ordinary hydrogen is replaced by deuterium) and is sensitive to the dissociation of the
deuterium nucleus by a high-energy neutrino. Similar techniques on a very large scale look for neutrinos passing
through large bodies of water or through the Antarctic ice sheets, such as the planned Ice Cube facility.
Meanwhile, physicists continue to work on ways to detect particles and waves that have never been directly observed
in the laboratory. An obvious example is provided by gravitational radiation, which is emitted by rapidly-accelerated
masses such as very close compact binary stars or objects falling into black holes. The LIGO observatory, currently in
the early stages of collecting data, consists of facilities in Louisiana and Washington, each with two evacuated arms

18

Cosmology Primer
four kilometers in length, arranged at right angles. Laser light is bounced off of mirrors suspended at the ends of each
arm, and sensitive detectors search for tiny displacements due to the stretching of spacetime characteristic of
gravitational waves. A future project is the LISA observatory, similar in spirit to the ground-based observatories but
consisting of three satellites flying in formation at distances of five million kilometers from each other. Like the
Planck mission to observe the CMB, LISA is a joint effort between NASA and the European Space Agency.
It goes without saying that we would like to directly observe the mysterious dark constituents of the universe: dark
matter and dark energy. Unfortunately there are many different models for what these might be, and correspondingly
many ways we might try to detect them (and the very real possibility that they might never be directly detectable). As
discussed in the page on the dark universe, the leading candidates for dark matter are either supersymmetric
neutralinos or axions; in both cases there are active programs underway to detect such particles directly. For
neutralinos, aside from finding evidence for supersymmetry at high-energy particle accelerators, the best bet is to
build extremely sensitive devices that would register each time a passing neutralino scattered off a nucleus in the
detector -- a very rare event indeed. To shield as best as one can from sources of external noise, such detectors are
typically built deep underground, so that the dirt above serves as a barrier to unwanted particles such as cosmic rays.
Axions, meanwhile, can be converted to photons in a precisely-tuned magnetic field, and experiments are underway to
search for such an effect. Unfortunately, the magnetic field must be tuned in a way that depends on the mass of the
axion, which is one of the thing we don't know with great precision.
For dark energy, meanwhile, there are three main possibilities: it is a strictly constant vacuum energy, a slowlyvarying dynamical component, or a breakdown of general relativity on cosmological scales. If the dark energy is
simply vacuum energy, there is no way to detect it directly; the best we can hope for is to attain a better understanding
of why it has the value it does, perhaps through hints provided by particle physics. If it is dynamical, however, it may
be possible to detect long-range forces due to gradual variations in the dark energy. Finally, if gravity is to blame, we
may be fortunate enough to detect a breakdown of general relativity in the Solar System due to the same effects that
are making it break down in cosmology; unfortunately, the specific effects we might see will depend on the details,
which are far from clear at this point.

19

Cosmology Primer

IX- Frequently Asked Questions

What is the universe expanding into?


Are distant galaxies moving faster than the speed of light? Wouldn't that violate relativity?
Does the universe have a center?
Could we detect the expansion of the universe by trying to measure the expansion of the solar system?
Is the universe finite or infinite? Will it recollapse or expand forever?
Is space flat or curved? I've heard both.
Is energy conserved in an expanding universe?
What is the difference between dark matter and dark energy?
Will we ever be able to detect dark matter or dark energy directly?
Isn't "dark energy" just like the older concept of the "ether"?
How do you know that dark matter isn't just ordinary matter that we can't see?
Could the inferred existence of dark matter and dark energy be due to a modified behavior of gravity?
Is inflation testable?
What came before the Big Bang?
Is our universe the only one, or are there others?

What is the universe expanding into?


As far as we know, the universe isn't expanding "into" anything. When we say the universe is expanding, we have a
very precise operational concept in mind: the amount of space in between distant galaxies is growing. (Individual
galaxies are not growing, as they are bound together by gravity.) But the universe is all there is (again, as far as we
know), so there's nothing outside into which it could be expanding. This is hard to visualize, since we are used to
thinking of objects as being located somewhere in space; but the universe includes all of space.

Are distant galaxies moving faster than the speed of light? Wouldn't that violate relativity?
A profound feature of relativity is that two objects passing by each other cannot have a relative velocity greater than
the speed of light. An even more profound feature, one which has received much less publicity, is that the concept of
"relative velocity" does not even make sense unless the objects are very close to each other. In Einstein's general
theory of relativity (which describes gravity as the curvature of spacetime), there is no way to define the velocity
between two widely-separated objects in any strictly correct sense. The "velocity" that cosmologists speak of between
distant galaxies is really just a shorthand for the expansion of the universe; it's not that the galaxies are moving, it's
that the space between them is expanding. If the distance isn't too great, this expansion looks and feels just like a
recession velocity, but when the distance becomes very large that resemblance breaks down. In particular, it's perfectly
plausible to have distant galaxies whose "recession velocity" is greater than the speed of light. (We couldn't see such
galaxies directly, since light from them would never reach us, but that doesn't mean they aren't there.) The resolution
to this paradox is simply that we have taken a convenient analogy too far, and there isn't a well-defined "speed"
between us and distant objects.

Does the universe have a center?


No. Our observable universe looks basically the same from the point of view of any observer. We see galaxies moving
away from us in all directions, but an astronomer living in any one of those galaxies would also see all the galaxies
(including our own) moving away from them. In particular, the Big Bang is not an explosion that happened at some
particular point in space; according to the Big Bang model, the entire universe came into existence expanding at every

20

Cosmology Primer
point all at once.

Could we detect the expansion of the universe by trying to measure the expansion of the solar
system?
No. Any system that is bound together by internal forces -- whether it is a table, the solar system, or the galaxy -- does
not expand along with the universe. (Not just that it only expands slightly; it really doesn't expand at all, or at least not
because of the expansion of the universe.) To observe the expansion, we need to study objects that are very distant, not
directly bound to us by gravity or anything else.

Is the universe finite or infinite? Will it recollapse or expand forever?


We don't really know in either case. Since the Big Bang happened a finite time ago (about 14 billion years), and since
light travels at a finite speed, there is an unbreakable upper limit to how far away we can see in the universe. Up to the
limits of the observable universe, what we observe is consistent with a uniform distribution of matter and energy that
could easily extend forever. On the other hand, it might eventually turn into something very different, beyond what we
can see; indeed, this might arise naturally as a result of inflation (see the really early universe). Similarly, we can
straightforwardly extrapolate the current evolution of our universe, dominated by dark energy, to predict a future in
which the universe continues to expand for all time (see the dark universe). However, the dark energy might someday
change its character into something different, in which case the universe might very well collapse. So, given how little
we currently understand about the nature of dark energy, we can't say anything for sure about the ultimate fate of our
universe.

Is space flat or curved? I've heard both.


There is an important distinction between "space" and "spacetime," and also a distinction between exact statements
and useful approximations. Our universe is a four-dimensional spacetime -- to describe the location of an event, you
need to specify three coordinates of space and one of time. According to Einstein, spacetime can be curved, and
gravitation is the manifestation of that spacetime curvature. Since there is certainly gravity in the universe, there is no
question that the universe is curved. But for cosmological purposes it is useful to model spacetime as a threedimensional space expanding as a function of time; then the total curvature is a combination of the curvature of space
by itself, plus the expansion of the universe. Observations indicate that space by itself is very nearly flat, rather than
having an overall positive or negative curvature (see the expanding universe); that is the origin of the statement that
we live in a "flat universe." Of course this is only an approximation, since the real world features galaxies and voids in
large-scale structure, rather than perfect smoothness; but it's a good approximation. So "space" is (approximately) flat,
while "spacetime" is definitely curved.

Is energy conserved in an expanding universe?


This is a tricky question, depending on what you mean by "energy." Usually we ascribe energy to the different
components of the universe (radiation, matter, dark energy), not including gravity itself. In that case the total energy,
given by adding up the energy density in each component, is certainly not conserved. The most dramatic example
occurs with dark energy -- the energy density (energy per unit volume) remains approximately constant, while the
volume increases as the universe expands, so the total energy increases. But even ordinary radiation exhibits similar
behavior; the number of photons remains constant, while each individual photon loses energy as it redshifts, so the
total energy in radiation decreases. (A decrease in energy is just as much a violation of energy conservation as an
increase would be.) In a sense, the energy in "stuff" is being transferred to the energy of the gravitational field, as
manifested in the expansion of the universe. But there is no exact definition of "the energy of the gravitational field,"
so this explanation is imperfect. Nevertheless, although energy is not really conserved in an expanding universe, there
is a very strict rule that is obeyed by the total energy, which reduces to perfect conservation when the expansion rate
goes to zero; the expansion changes the rules, but that doesn't mean that anything goes.

What is the difference between dark matter and dark energy?


Dark matter behaves much like a collection of ordinary matter made of particles, except that it's dark. In particular,
dense regions of dark matter tend to become even more dense, as the mutual gravitational force of the matter pulls it

21

Cosmology Primer
together. For this reason, we suspect that the dark matter is some sort of new, massive particle, just one we haven't yet
discovered in the laboratory (yet). Dark energy, on the other hand, doesn't act anything like particles: it doesn't cluster
together, nor does it dilute as the universe expands. Its density remains constant (so far as we can tell) throughout
space and time. So whatever the dark energy is, it's something different than dark matter.

Will we ever be able to detect dark matter or dark energy directly?


Hopefully. Different candidates for what the dark matter particles are lead to different strategies for detecting them,
either directly in laboratories here on Earth or indirectly through high-energy particles from space. But numerous
efforts are being undertaken, and we might find dark matter in the near future. Dark energy is an even longer shot; if it
is a strictly constant vacuum energy, we could never detect it directly, while a dynamical field could conceivably be
detected. Probably we will have to content ourselves with understanding dark energy indirectly, through its
gravitational effects on the expansion of the universe. See the page on the dark universe.

Isn't "dark energy" just like the older concept of the "ether"?
No; in fact, it's just the opposite. The ether was supposed to be an invisible substance that determined the rest frame of
the universe. It was expected by theorists, but eventually abandoned when experimenters could not find any evidence
for it (and Einstein figured out that it wasn't necessary). Dark energy, meanwhile, was not at all expected by most
working cosmologists; we need it to explain observed facts, like the acceleration of the universe and the mismatch
between matter and total energy. And the dark energy appears the same to all observers, so there's no sense in which it
determines a rest frame.

How do you know that dark matter isn't just ordinary matter that we can't see?
We can measure the total amount of matter through the gravitational field it creates, both in galaxies and in clusters of
galaxies. But we can separately measure the amount of ordinary matter by less direct means. The traditional method is
to study the abundances of light elements (hydrogen, deuterium, helium, and lithium) in the early universe. These
elements are produced by primordial nucleosynthesis (see the early universe), and the amount of ordinary matter in
the universe directly affects the relative amounts of different elements that we predict. Observations of these elements
are consistent with ordinary matter comprising only 5% of the total energy of the universe, whereas the total amount
of matter is closer to 30%, with dark matter making up the difference. Confirmation of this result comes from
temperature fluctuations in the cosmic microwave background; the precise pattern of these fluctuations depends on the
ratio of ordinary matter to dark matter in a way which matches the requirements of primordial nucleosynthesis. So
there are strong (and independent) reasons to believe that the dark matter is something new, not just ordinary matter
that is somehow hiding.

Could the inferred existence of dark matter and dark energy be due to a modified behavior of
gravity?
It's possible, and in fact there are scientists working hard on just this scenario -- doing away with dark matter and/or
dark energy, and instead invoking a new law of gravity on very large scales. There are a few obstacles to this idea,
though, and two are worth stressing. One is that the dark matter/dark energy paradigm does an extremely good job of
explaining the data, and in a wide variety of apparently disconnected circumstances. The other is that our current
theory of gravity -- Einstein's general theory of relativity -- is both conceptually compelling and experimentally very
well tested. It's hard to come up with a new theory that fits the data nearly as well as the conventional model of dark
matter and dark energy in the framework of general relativity (but that's no reason not to keep trying).

Is inflation testable?
Yes and no. Inflation makes quite strong predictions, including a geometrically flat universe (already verified by
measurements of the cosmic microwave background) and a particular set of primordial perturbations, both
fluctuations in the matter density and in gravitational waves. The fluctuations in the matter density seem consistent
with the predictions of inflation, while looking for gravitational waves from inflation is a major goal of experimenters.
However, there are two caveats. First, there are many different models of inflation, and they give somewhat different
predictions, so it's possible to wriggle out of almost any definitive statement. And second, the predictions of inflation

22

Cosmology Primer
may also be predictions of some alternative model which has not yet been thought of. We will never prove inflation
beyond any possible doubt; we will only gain increasing (or decreasing) confidence in the inflationary paradigm, as
developments in both theory and experiment either remain consistent with inflation or make it seem less likely.

What came before the Big Bang?


The strictly correct answer is: nobody knows, and nobody even knows if the question makes sense. According to
general relativity, Einstein's theory of gravity and our best understanding of what governs the early universe, there is
no such thing as "before the Big Bang" -- it is the point at which space and time come into existence. However, it is
also a "singular" point, at which our theories break down. It is possible that some future reconciliation of general
relativity with quantum mechanics will help us understand the origin of the Big Bang, just as it is possible that we
may come to believe that the universe had an interesting history even before what we now call the Bang. Both
possibilities are being actively pursued by cosmologists.

Is our universe the only one, or are there others?


Hopefully you won't be disappointed if we say that we don't know. There are different kinds of "other universes" that
one could reasonably imagine -- other regions of space that are very far away and look very different, or regions that
are separated from our own by extra dimension of space, or different branches of the quantum-mechanical
wavefunction of the universe. These are all profound ideas which we won't discuss in detail here. Suffice it to say that
these kinds of other universes are perfectly plausible, and are sometimes even predicted by ambitious theories of
fundamental physics. However, it is hard to see how we could test their existence experimentally. So we don't know
one way or the other, but speculations along these lines play an important role in the attempt to construct a unified
framework of physics and cosmology; perhaps in the future we will be able to be more definite.

23

Das könnte Ihnen auch gefallen