A: The electromagnetic radiation field
In this appendix, we will briefly review the most
important properties of a radiation field. We thereby assume that
the reader has encountered these quantities already in a different
context.
A.1 Parameters of the radiation field
The electromagnetic radiation field is described
by the specific intensity I
ν , which is
defined as follows. Consider a surface element of area d
A . The radiation energy
which passes through this area per time interval d t from within a solid angle element d
ω around a direction
described by the unit vector
, with frequency in the range
between ν and ν + d ν , is
where θ describes the angle
between the direction
of the light and the normal
vector of the surface element. Then, d A cos θ is the area projected in the
direction of the infalling light. The specific intensity depends on
the considered position (and, in time-dependent radiation fields,
on time), the direction
, and the frequency ν . With the definition ( A.1 ), the dimension of
I ν is energy per unit area, time,
solid angle, and frequency, and it is typically measured in units
of
.
The specific intensity of a cosmic source describes its surface
brightness.


(A.1)



The specific net
flux F ν
passing through an area element is obtained by integrating the
specific intensity over all solid angles,
The flux that we receive from a cosmic source is defined in exactly
the same way, except that cosmic sources usually subtend a very
small solid angle on the sky. In calculating the flux we receive
from them, we may therefore drop the factor cos θ in ( A.2 ); in this context,
the specific flux is also denoted as S ν . However, in this Appendix
(and only here!), the notation S ν will be reserved for another
quantity. The flux is measured in units of
. If the
radiation field is isotropic, F ν vanishes. In this case, the
same amount of radiation passes through the surface element in both
directions.

(A.2)

The mean
specific intensity J ν is defined as the average of
I ν over all angles,
so that, for an isotropic radiation field, I ν = J ν . The specific energy density u
ν is related to
J ν according to
where u ν is the energy of the radiation
field per volume element and frequency interval, thus measured in
. The total energy
density of the radiation is obtained by integrating u ν over frequency. In the same
way, the intensity of the radiation is obtained by integrating the
specific intensity I
ν over
ν .

(A.3)

(A.4)

A.2 Radiative transfer
The specific intensity of radiation in the
direction of propagation between source and observer is constant,
as long as no emission or absorption processes are occurring. If
s measures the length along
a line-of-sight, the above statement can be formulated as
An immediate consequence of this equation is that the surface
brightness of a source is independent of its distance. The observed
flux of a source depends on its distance, because the solid angle,
under which the source is observed, decreases with the square of
the distance, F
ν ∝
D −2 [see (
A.2 )].
However, for light propagating through a medium, emission and
absorption (or scattering of light) occurring along the path over
which the light travels may change the specific intensity. These
effects are described by the equation of radiative transfer
The first term describes the absorption of radiation and states
that the radiation absorbed within a length interval d s is proportional to the incident
radiation. The factor of proportionality is the absorption coefficient κ
ν , which has
the unit of cm −1 . The emission coefficient j
ν describes the
energy that is added to the radiation field by emission processes,
having a unit of
; hence, it is the radiation energy emitted per volume element,
time interval, frequency interval, and solid angle. Both,
κ ν and j ν depend on the nature and state
(such as temperature, chemical composition) of the medium through
which light propagates.

(A.5)

(A.6)

The absorption and emission coefficients both
account for true absorption and emission processes, as well as the
scattering of radiation. Indeed, the scattering of a photon can be
considered as an absorption that is immediately followed by an
emission of a photon.
The optical
depth τ ν
along a line-of-sight is defined as the integral over the
absorption coefficient,
where s 0
denotes a reference point on the sightline from which the optical
depth is measured. Dividing ( A.6 ) by κ ν and using the relation d
τ ν = κ ν d s in order to introduce the optical
depth as a new variable along the light ray, the equation of
radiative transfer can be written as
where the source function
is defined as the ratio of the emission and absorption
coefficients. In this form, the equation of radiative transport can
be formally solved; as can easily be tested by substitution, the
solution is
This equation has a simple interpretation. If I ν (0) is the incident intensity,
it will have decreased by absorption to a value
after
an optical depth of τ
ν . On the other
hand, energy is added to the radiation field by emission, accounted
for by the τ
′ -integral.
Only a fraction
of this additional energy emitted at τ ′ reaches the point τ , the rest is absorbed.

(A.7)

(A.8)

(A.9)

(A.10)


In the context of ( A.10 ), we call this a
formal solution for the
equation of radiative transport. The reason for this is based on
the fact that both the absorption and the emission coefficient
depend on the physical state of the matter through which radiation
propagates, and in many situations this state depends on the
radiation field itself. For instance, κ ν and j ν depend on the temperature of
the matter, which in turn depends, by heating and cooling
processes, on the radiation field it is exposed to. Hence, one
needs to solve a coupled system of equations in general: on the one
hand the equation of radiative transport, and on the other hand the
equation of state for matter. In many situations, very complex
problems arise from this, but we will not consider them further in
the context of this book.
A.3 Blackbody radiation
For matter in thermal equilibrium, the source
function S
ν is solely a
function of the matter temperature,
independent of the composition of the medium (Kirchhoff’s law). We
will now consider radiation propagating through matter in thermal
equilibrium at constant temperature T . Since in this case S ν = B ν ( T ) is constant, the solution (
A.10 ) can be
written in the form
From this it follows that I
ν =
B ν ( T ) is valid for sufficiently large
optical depth τ
ν . The
radiation propagating through matter which is in thermal
equilibrium is described by the function B ν ( T ) if the optical depth is
sufficiently large, independent of the composition of the matter. A
specific case of this situation can be illustrated by imagining the
radiation field inside a box whose opaque walls are kept at a
constant temperature T .
Due to the opaqueness of the walls, their optical depth is
infinite, hence the radiation field within the box is given by
I ν = B ν ( T ). This is also valid if the volume
is filled with matter, as long as the latter is in thermal
equilibrium at temperature T . For these reasons, this kind of
radiation field is also called blackbody radiation.

(A.11)
![$$\displaystyle\begin{array}{rcl} I_{\nu }(\tau _{\nu })& =& I_{\nu }(0)\,\exp \left (-\tau _{\nu }\right ) \\ & & +B_{\nu }(T)\int _{0}^{\tau _{\nu }}\mathrm{d}\tau _{\nu }^{{\prime}}\;\exp \left (\tau _{\nu }^{{\prime}}-\tau _{\nu }\right ) \\ & =& I_{\nu }(0)\,\exp \left (-\tau _{\nu }\right ) + B_{\nu }(T)\,\left [1 -\exp \left (-\tau _{\nu }\right )\right ]\;.{}\end{array}$$](A129044_2_En_BookBackmatter_OnlinePDF_Equ12.gif)
(A.12)
The function B ν ( T ) was first obtained in 1900 by Max
Planck, and in his honor, it was named the Planck function ; it reads
where
is the Planck constant and
is the Boltzmann constant. The shape of the spectrum (Fig.
A.1 ) can be
derived from statistical physics. Blackbody radiation is defined by
I ν = B ν ( T ), and thermal radiation by S ν = B ν ( T ). For large optical depths in the
case of thermal radiation, the specific intensity converges to
blackbody radiation. For small optical depth, the radiation field
is approximated by an integral over the emissivity j ν , which can deviate strongly
from that of blackbody spectrum even in the case of a thermal
source; an example is the optically thin bremsstrahlung from the
hot gas in galaxy clusters (see Sect. 6.4 ).

(A.13)



Fig. A.1
The Planck function ( A.13 ) for different
temperatures T . The plot
shows B ν ( T ) as a function of frequency
ν , where high frequencies
are plotted towards the left (thus large wavelengths towards the
right). The exponentially decreasing Wien part of the spectrum is
visible on the left , the
Rayleigh–Jeans part on the right . The shape of the spectrum in the
Rayleigh–Jeans part is independent of the temperature, which is
determining the amplitude however. Credit: T. Kaempf & M.
Altmann, Argelander-Institut für Astronomie, Universität Bonn
The Planck function has its maximum at
i.e., the frequency of the maximum is proportional to the
temperature. This property is called Wien’s law . This law can also be
written in more convenient units,

(A.14)

(A.15)
The Planck function can also be formulated
depending on wavelength
, such that B λ ( T ) d λ = B ν ( T ) d ν ,
Two limiting cases of the Planck function are of particular
interest. For low frequencies, h P ν ≪ k B T , one can apply the expansion of the
exponential function for small arguments in ( A.13 ). The leading-order
term in this expansion then yields
which is called the Rayleigh–Jeans
approximation of the Planck function. We point out that the
Rayleigh–Jeans equation does not contain the Planck constant, and
this law had been known even before Planck derived his exact
equation. In the other limiting case of very high frequencies,
h P ν ≫ k B T , the exponential factor in the
denominator in ( A.13 ) becomes very much larger than unity, so
that we obtain
called the Wien
approximation of the Planck function.


(A.16)

(A.17)

(A.18)
The energy density of blackbody radiation depends
only on the temperature, of course, and is calculated by
integration over the Planck function,
where we defined the frequency-integrated Planck function
and where the constant a
has the value
The flux which is emitted by the surface of a blackbody per unit
area is given by
where the Stefan–Boltzmann
constant σ SB has a value of

(A.19)

(A.20)

(A.21)

(A.22)

(A.23)
A.4 The magnitude scale
Optical astronomy was being conducted well before
methods of quantitative measurements became available. The
brightness of stars had been cataloged more than 2000 years ago,
and their observation goes back as far as the ancient world. Stars
were classified into magnitudes, assigning a magnitude of 1 to the
brightest stars and higher magnitudes to the fainter ones. Since
the apparent magnitude as perceived by the human eye scales roughly
logarithmically with the radiation flux (which is also the case for
our hearing), the magnitude scale represents a logarithmic flux
scale. To link these visually determined magnitudes in historical
catalogs to a quantitative measure, the magnitude system has been
retained in optical astronomy, although with a precise definition.
Since no historical astronomical observations have been conducted
in other wavelength ranges, because these are not accessible to the
unaided eye, only optical astronomy has to bear the historical
burden of the magnitude system.
A.4.1 Apparent magnitude
We start with a relative system of flux
measurements by considering two sources with fluxes S 1 and S 2 . The apparent magnitudes of the two sources,
m 1 and
m 2 , then
behave according to
This means that the brighter source has a smaller apparent
magnitude than the fainter one: the larger the apparent magnitude,
the fainter the source. 1 The factor of 2.5 in this definition
is chosen so as to yield the best agreement of the magnitude system
with the visually determined magnitudes. A difference of |
Δ m | = 1 in this system
corresponds to a flux ratio of ∼ 2. 51, and a flux ratio of a
factor 10 or 100 corresponds to 2.5 or 5 magnitudes,
respectively.

(A.24)
A.4.2 Filters and colors
Since optical observations are performed using a
combination of a filter and a detector system, and since the flux
ratios depend, in general, on the choice of the filter (because the
spectral energy distribution of the sources may be different),
apparent magnitudes are defined for each of these filters. The most
common filters are shown in Fig. A.2 and listed in
Table A.1 , together with their characteristic
wavelengths and the widths of their transmission curves. The
apparent magnitude for a filter X is defined as m X , frequently written as
X . Hence, for the B-band
filter, m B ≡ B .

Fig. A.2
Transmission curves of various
filter-detector systems. From top
to bottom : the filters of the NICMOS camera and the WFPC2
on-board HST, the Washington filter system, the filters of the EMMI
instrument at ESO’s NTT, the filters of the WFI at the ESO/MPG
2.2-m telescope and those of the SOFI instrument at the NTT, and
the Johnson-Cousins filters. In the bottom diagram, the spectra of three
stars with different effective temperatures are displayed. Source:
L. Girardi et al. 2002, Theoretical isochrones in several photometric
systems. I. Johnson-Cousins-Glass, HST/WFPC2, HST/NICMOS,
Washington, and ESO Imaging Survey filter sets , A&A
391, 195, p. 204, Fig. 3. ©ESO. Reproduced with
permission
Next, we need to specify how the magnitudes
measured in different filters are related to each other, in order
to define the color indices of sources. For this purpose, a
particular class of stars is used, main-sequence stars of spectral
type A0, of which the star Vega is an archetype. For such a star,
by definition,
, i.e., every color index
for such a star is defined to be zero.

For a more precise definition, let T X ( ν ) be the transmission curve of the
filter-detector system. T
X ( ν ) specifies which fraction of the
incoming photons with frequency ν are registered by the detector. The
apparent magnitude of a source with spectral flux S ν is then
where the constant needs to be determined from reference
stars.

(A.25)
Another commonly used definition of magnitudes is
the AB system. In contrast to the Vega magnitudes, no stellar
spectral energy distribution is used as a reference here, but
instead one with a constant flux at all frequencies,
. This value has been chosen such that A0 stars like Vega have the
same magnitude in the original Johnson-V-band as they have in the
AB system, m
V AB
= m V . With ( A.25 ), one obtains for
the conversion between the two systems
For the filters at the ESO Wide-Field Imager, which are designed to
resemble the Johnson set of filters, the following prescriptions
are then to be applied:
;
;
V AB =
V Vega ;
;
.


(A.26)




Table A.1
For some of the best-established filter
systems—Johnson, Strömgren, and the filters of the Sloan Digital
Sky Surveys—the central (more precisely, the effective) wavelengths
and the widths of the filters are listed
![]() |
|
![]() |
|
![]() |
A.4.3 Absolute magnitude
The apparent magnitude of a source does not in
itself tell us anything about its luminosity, since for the
determination of the latter we also need to know its distance
D in addition to the
radiative flux. Let L
ν be the
specific luminosity of a source, i.e., the energy emitted per unit
time and per unit frequency interval, then the flux is given by
(note that from here on we switch back to the notation where
S denotes the flux, which
was denoted by F earlier in
this appendix)
where we implicitly assumed that the source emits isotropically.
Having the apparent magnitude as a measure of S ν (at the frequency ν defined by the filter which is
applied), it is desirable to have a similar measure for
L ν , specifying the physical
properties of the source itself. For this purpose, the absolute magnitude is introduced,
denoted as M
X , where X
refers to the filter under consideration. By definition,
M X is equal to the apparent
magnitude of a source if it were to be located at a distance of
10 pc from us. The absolute magnitude of a source is thus
independent of its distance, in contrast to the apparent magnitude.
With ( A.27 )
we find for the relation of apparent to absolute magnitude
where we have defined the distance
modulus μ in the final step. Hence, the latter is a
logarithmic measure of the distance of a source: μ = 0 for D = 10 pc, μ = 10 for D = 1 kpc, and μ = 25 for D = 1 Mpc. The difference between
apparent and absolute magnitude is independent of the filter
choice, and it equals the distance modulus if no extinction is
present. In general, this difference is modified by the wavelength-
(and thus filter-)dependent extinction coefficient—see Sect.
2.2.4 .

(A.27)

(A.28)
A.4.4 Bolometric parameters
The total luminosity L of a source is the integral of the
specific luminosity L
ν over all
frequencies. Accordingly, the total flux S of a source is the
frequency-integrated specific flux S ν . The apparent bolometric magnitude m
bol is defined as a logarithmic measure of the total
flux,
where here the constant is also determined from reference stars.
Accordingly, the absolute
bolometric magnitude is defined by means of the distance
modulus, as in ( A.28 ). The absolute bolometric magnitude
depends on the bolometric luminosity L of a source via
The constant can be fixed, e.g., by using the parameters of the
Sun: its apparent bolometric magnitude is
, and the distance of
one Astronomical Unit corresponds to a distance modulus of
. With these values, the absolute
bolometric magnitude of the Sun becomes
so that ( A.30 ) can be written as
and the luminosity of the Sun is then

(A.29)

(A.30)



(A.31)

(A.32)

(A.33)
The direct relation between bolometric magnitude
and luminosity of a source can hardly be exploited in practice,
because the apparent bolometric magnitude (or the flux S ) of a source cannot be observed in
most cases. For observations of a source from the ground, only a
limited window of frequencies is accessible. Nevertheless, in these
cases one also likes to quantify the total luminosity of a source.
For sources for which the spectrum is assumed to be known, like for
many stars, the flux from observations at optical wavelengths can
be extrapolated to larger and smaller wavelengths, and so
m bol can be
estimated. For galaxies or AGNs, which have a much broader spectral
distribution and which show much more variation between the
different objects, this is not feasible. In these cases, the flux
of a source in a particular frequency range is compared to the flux
the Sun would have at the same distance and in the same spectral
range. If M
X is the
absolute magnitude of a source measured in the filter X, the X-band
luminosity of this source is defined as
Thus, when speaking of, say, the ‘blue luminosity of a galaxy’,
this is to be understood as defined in ( A.34 ). For reference,
the absolute magnitude of the Sun in optical filters is
M ⊙U = 5. 55,
M ⊙B = 5. 45,
M ⊙V = 4. 78,
M ⊙R = 4. 41,
and M ⊙I
= 4. 07.

(A.34)
B: Properties of stars
In this appendix, we will summarize the most
important properties of stars as they are required for
understanding the contents of this book. Of course, this brief
overview cannot replace the study of other textbooks in which the
physics of stars is covered in much more detail.
B.1 The parameters of stars
To a good approximation, stars are gas spheres,
in the cores of which light atomic nuclei are transformed into
heavier ones (mainly hydrogen into helium) by thermonuclear
processes, thereby producing energy. The external appearance of a
star is predominantly characterized by its radius R and its characteristic temperature
T . The properties and
evolution of a star depend mainly on its mass M .
In a first approximation, the spectral energy
distribution of the emission from a star can be described by a
blackbody spectrum. This means that the specific intensity
I ν is given by a Planck spectrum (
A.13 ) in
this approximation. The luminosity L of a star is the energy radiated per
unit time. If the spectrum of star was described by a Planck
spectrum, the luminosity would depend on the temperature and on the
radius according to
where ( A.22
) was applied. However, the spectra of stars deviate from that of a
blackbody (see Fig. 3.33 and the bottom panel of Fig.
A.2 ). One
defines the effective temperature
T eff of a star as the temperature a blackbody of
the same radius would need to have to emit the same luminosity as
the star, thus
The luminosities of stars cover a huge range; the weakest are a
factor ∼ 10 4 times less luminous than the Sun, whereas
the brightest emit ∼ 10 5 times as much energy per unit
time as the Sun. This big difference in luminosity is caused either
by a variation in radius or by different temperatures. We know from
the colors of stars that they have different temperatures: there
are blue stars which are considerably hotter than the Sun, and red
stars that are very much cooler. The temperature of a star can be
estimated from its color. From the flux ratio at two different
wavelengths or, equivalently, from the color index
in two filters X and Y,
the temperature T
c is determined such that a blackbody at T c would have the same
color index. T c
is called the color
temperature of a star. If the spectrum of a star was a
Planck spectrum, then the equality T c = T eff would hold, but in
general these two temperatures differ.

(B.1)

(B.2)

B.2 Spectral class, luminosity class, and the Hertzsprung–Russell diagram
The spectra of stars can be classified according
to the atomic (and, in cool stars, also molecular) spectral lines
that are present. Based on the line strengths and their ratios, the
Harvard sequence of stellar spectra was introduced. These spectral
classes follow a sequence that is denoted by the letters O, B, A,
F, G, K, M; besides these, some other spectral classes exist that
will not be mentioned here. The sequence corresponds to a sequence
of color temperature of stars: O stars are particularly hot, around
50 000 K, M stars very much cooler with T c ∼ 3500 K. For a finer
classification, each spectral class is supplemented by a number
between 0 and 9. An A1 star has a spectrum very similar to that of
an A0 star, whereas an A5 star has as many features in common with
an A0 star as with an F0 star.

Fig. B.1
Color-magnitude diagram for 41 453
individual stars, whose parallaxes were determined by the Hipparcos
satellite with an accuracy of better than 20 %. Since the stars
shown here are subject to unavoidable strong selection effects
favoring nearby and luminous stars, the relative number density of
stars is not representative of their true abundance. In particular,
the lower main sequence is much more densely populated than is
visible in this diagram. Credit: European Space Agency, Web page of
the Hipparcos project
Plotting the spectral type versus the absolute
magnitude for those stars for which the distance and hence the
absolute magnitude can be determined, a striking distribution of
stars becomes apparent in such a Hertzsprung–Russell diagram (HRD).
Instead of the spectral class, one may also plot the color index of
the stars, typically B −
V or V − I . The resulting color-magnitude diagram (CMD) is
essentially equivalent to an HRD, but is based solely on
photometric data. A different but very similar diagram plots the
luminosity versus the effective temperature.
In Fig. B.1 , a color-magnitude diagram is plotted,
compiled from data observed by the HIPPARCOS satellite. Instead of
filling the two-dimensional parameter space rather uniformly,
characteristic regions exist in such color-magnitude diagrams in
which nearly all stars are located. Most stars can be found in a
thin band called the main
sequence . It extends from early spectral types (O, B) with
high luminosities (‘top left’) down to late spectral types (K, M)
with low luminosities (‘bottom right’). Branching off from this
main sequence towards the ‘top right’ is the domain of red giants,
and below the main sequence, at early spectral types and very much
lower luminosities than on the main sequence itself, we have the
domain of white dwarfs. The fact that most stars are arranged along
a one-dimensional sequence—the main sequence—is probably one of the
most important discoveries in astronomy, because it tells us that
the properties of stars are determined basically by a single
parameter: their mass.

Fig. B.2
Schematic color-magnitude diagram in which
the spectral types and luminosity classes are indicated. Source:
http://de.wikipedia.org
Since stars exist which have, for the same
spectral type and hence the same color temperature (and roughly the
same effective temperature), very different luminosities, we can
deduce immediately that these stars have different radii, as can be
read from ( B.2 ). Therefore, stars on the red giant
branch, with their much higher luminosities compared to
main-sequence stars of the same spectral class, have a much larger
radius than the corresponding main-sequence stars. This size effect
is also observed spectroscopically: the gravitational acceleration
on the surface of a star (surface gravity) is
We know from models of stellar atmospheres that the width of
spectral lines depends on the gravitational acceleration on the
star’s surface: the lower the surface gravity, the narrower the
stellar absorption lines. Hence, a relation exists between the line
width and the stellar radius. Since the radius of a star—for a
fixed spectral type or effective temperature—specifies the
luminosity, this luminosity can be derived from the width of the
lines. In order to calibrate this relation, stars of known distance
are required.

(B.3)
Based on the width of spectral lines, stars are
classified into luminosity
classes : stars of luminosity class I are called
supergiants, those of luminosity class III are giants,
main-sequence stars are denoted as dwarfs and belong to luminosity
class V; in addition, the classification can be further broken
down into bright giants (II), subgiants (IV), and subdwarfs (VI).
Any star in the Hertzsprung–Russell diagram can be assigned a
luminosity class and a spectral class (Fig. B.2 ). The Sun is a
G2 star of luminosity class V.
If the distance of a star, and thus its
luminosity, is known, and if in addition its surface gravity can be
derived from the line width, we obtain the stellar mass from these
parameters. By doing so, it turns out that for main-sequence stars
the luminosity is a steep function of the stellar mass,
approximately described by
Therefore, a main-sequence star of M = 10 M ⊙ is ∼ 3000 times more
luminous than our Sun.

(B.4)
B.3 Structure and evolution of stars
To a very good approximation, stars are
spherically symmetric. Therefore, the structure of a star is
described by the radial profile of the parameters of its stellar
plasma. These are density, pressure, temperature, and chemical
composition of the matter. During almost the full lifetime of a
star, the plasma is in hydrostatic equilibrium, so that pressure
forces and gravitational forces are of equal magnitude and directed
in opposite directions, so as to balance each other.
The density and temperature are sufficiently high
in the center of a star that thermonuclear reactions are ignited.
In main-sequence stars, hydrogen is fused into helium, thus four
protons are combined into one 4 He nucleus. For every
helium nucleus that is produced this way, 26. 73 MeV of energy are
released. Part of this energy is emitted in the form of neutrinos
which can escape unobstructed from the star due to their very low
cross section. 2 The
energy production rate is approximately proportional to
T 4 for
temperatures below about 15 × 10 6 K, at which the
reaction follows the so-called pp-chain. At higher temperatures,
another reaction chain starts to contribute, the so-called CNO
cycle, with an energy production rate which is much more strongly
dependent on temperature—roughly proportional to T 20 .
The energy generated in the interior of a star is
transported outwards, where it is then released in the form of
electromagnetic radiation. This energy transport may take place in
two different ways: first, by radiation transport, and second, it
can be transported by macroscopic flows of the stellar plasma. This
second mechanism of energy transport is called convection; here,
hot elements of the gas rise upwards, driven by buoyancy, and at
the same time cool ones sink downwards. The process is similar to
that observed in heating water on a stove. Which of the two
processes is responsible for the energy transport depends on the
temperature profile inside the star. The intervals in a star’s
radius in which energy transport takes place via convection are
called convection zones. Since in convection zones stellar material
is subject to mixing, the chemical composition is homogeneous
there. In particular, chemical elements produced by nuclear fusion
are transported through the star by convection.

Fig. B.3
Theoretical temperature-luminosity diagram
of stars. The solid curve
is the zero age main sequence (ZAMS), on which stars ignite the
burning of hydrogen in their cores. The evolutionary tracks of
these stars are indicated by the various lines which are labeled with the
stellar mass. The hatched
areas mark phases in which the evolution proceeds only
slowly, so that many stars are observed to be in these areas.
Source: A. Maeder & G. Meynet 1989, Grids of evolutionary models from 0.85 to 120
solar masses - Observational tests and the mass limits ,
A&A 210, 155, p. 166, Fig. 15. ©ESO. Reproduced with
permission
Stars begin their lives with a homogeneous
chemical composition, resulting from the composition of the
molecular cloud out of which they are formed. If their mass exceeds
about 0. 08 M ⊙
, the temperature and pressure in their core are sufficient to
ignite the fusion of hydrogen into helium. Gas spheres with a mass
below ∼ 0. 08 M
⊙ will not satisfy these conditions, hence these
objects—they are called brown dwarfs—are not stars in a proper
sense. 3 At the onset
of nuclear fusion, the star is located on the zero-age main
sequence (ZAMS) in the HRD (see Fig. B.3 ). The energy
production by fusion of hydrogen into helium alters the chemical
composition in the stellar interior; the abundance of hydrogen
decreases by the same rate as the abundance of helium increases. As
a consequence, the duration of this phase of central hydrogen
burning is limited. As a rough estimate, the conditions in a star
will change noticeably when about 10 % of its hydrogen is used up.
Based on this criterion, the lifetime of a star on the main
sequence can now be estimated. The total energy produced in this
phase can be written as
where M c 2 is
the rest-mass energy of the star, of which a fraction of 0.1 is
fused into helium, which is supposed to occur with an efficiency of
0.007. Phrased differently, in the fusion of four protons into one
helium nucleus, an energy of ∼ 0. 007 × 4 m p c 2 is generated, with
m p denoting the
proton mass. In particular, ( B.5 ) states that the total energy produced
during this main-sequence phase is proportional to the mass of the
star. In addition, we know from ( B.4 ) that the luminosity
is a steep function of the stellar mass. The lifetime of a star on
the main sequence can then be estimated by equating the available
energy E MS with
the product of luminosity and lifetime. This yields
Using this argument, we observe that stars of higher mass conclude
their lives on the main sequence much faster than stars of lower
mass. The Sun will remain on the main sequence for about eight to
ten billion years, with about half of this time being over already.
In comparison, very luminous stars, like O and B stars, will
have a lifetime on the main sequence of only a few million years
before they have exhausted their hydrogen fuel.

(B.5)

(B.6)
In the course of their evolution on the main
sequence, stars move away only slightly from the ZAMS in the HRD,
towards somewhat higher luminosities and lower effective
temperatures. In addition, the massive stars in particular can lose
part of their initial mass by stellar winds. The evolution after
the main-sequence phase depends on the stellar mass. Stars of very
low mass, M ≲ 0. 7
M ⊙ , have a
lifetime on the main sequence which is longer than the age of the
Universe, therefore they cannot have moved away from the main
sequence yet.
For massive stars, M ≳ 2. 5 M ⊙ , central hydrogen
burning is first followed by a relatively brief phase in which the
fusion of hydrogen into helium takes place in a shell outside the
center of the star. During this phase, the star quickly moves to
the ‘right’ in the HRD, towards lower temperatures, and thereby
expands strongly. After this phase, the density and temperature in
the center rise so much as to ignite the fusion of helium into
carbon. A central helium-burning zone will then establish itself,
in addition to the source in the shell where hydrogen is burned. As
soon as the helium in the core has been exhausted, a second shell
source will form fusing helium. In this stage, the star will become
a red giant or supergiant, ejecting part of its mass into the
interstellar medium in the form of stellar winds. Its subsequent
evolutionary path depends on this mass loss. A star with an initial
mass M ≲ 8 M ⊙ will evolve into a white
dwarf, which will be discussed further below.
For stars with initial mass M ≲ 2. 5 M ⊙ , the helium burning in
the core occurs explosively, in a so-called helium flash. A large
fraction of the stellar mass is ejected in the course of this
flash, after which a new stable equilibrium configuration is
established, with a helium shell source burning beside the
hydrogen-burning shell. Expanding its radius, the star will evolve
into a red giant or supergiant and move along the asymptotic giant
branch (AGB) in the HRD.
The configuration in the helium shell source is
unstable, so that its burning will occur in the form of pulses.
After some time, this will lead to the ejection of the outer
envelope which then becomes visible as a planetary nebula . The remaining
central star moves to the left in the HRD, i.e., its temperature
rises considerably (to more than 10 5 K). Finally, its
radius gets smaller by several orders of magnitude, so that the
stars move downwards in the HRD, thereby slightly reducing its
temperature: a white dwarf is born, with a mass of about 0. 6
M ⊙ and a radius
roughly corresponding to that of the Earth.
If the initial mass of the star is ≳ 8
M ⊙ , the
temperature and density at its center become so large that carbon
can also be fused. Subsequent stellar evolution towards a
core-collapse supernova is described in Sect. 2.3.2 .
The individual phases of stellar evolution have
very different time-scales. As a consequence, stars pass through
certain regions in the HRD very quickly, and for this reason stars
at those evolutionary stages are never or only rarely found in the
HRD. By contrast, long-lasting evolutionary stages like the main
sequence or the red giant branch exist, with those regions in an
observed HRD being populated by numerous stars.
C: Units and constants
In this book, we consistently used, besides
astronomical units, the Gaussian cgs system of units, with lengths
measured in cm, masses in g, and energies in erg. This is the
commonly used system of units in astronomy. In these units, the
speed of light is
, the
masses of protons, neutrons, and electrons are
,
,
and
,
respectively.




Frequently used units of length in astronomy
include the Astronomical Unit, thus the average separation between
the Earth and the Sun, where 1 AU = 1. 496 × 10 13 cm,
and the parsec (see Sect. 2.2.1 for the definition),
1 pc = 3. 086 × 10 18 cm. A year has 1 yr = 3. 156 × 10
7 s. In addition, masses are typically specified in
Solar masses, 1 M
⊙ = 1. 989 × 10 33 g, and the bolometric
luminosity of the Sun is
.

In cgs units, the value of the elementary charge
is
, and the unit of the magnetic field strength is one Gauss, where
. One of the very convenient properties of cgs units is that the
energy density of the magnetic field in these units is given by
—the reader may check that
the units of this equation are consistent.



X-ray astronomers measure energies in electron
Volts, where
.
Temperatures can also be measured in units of energy, because
k B T has the dimension of energy. They are
related according to 1 eV = 1. 161 × 10 4 k B K. Since we always use
the Boltzmann constant k
B in combination with a temperature, its actual value is
never needed. The same holds for Newton’s constant of gravity which
is always used in combination with a mass. Here one has
which can also be written in the form


(C.1)

(C.2)
The frequency of a photon is linked to its energy
according to h P
ν = E , and we have the relation
. Accordingly, we can write the wavelength
in the form



D: Recommended literature
In the following, we will give some
recommendations for further study of the literature on
astrophysics. For readers who have been in touch with astronomy
only occasionally until now, the general textbooks may be of
particular interest. The choice of literature presented here is a
very subjective one which represents the preferences of the author,
and of course it represents only a small selection of the many
astronomy texts available.
D.1 General textbooks
There exist a large selection of general
textbooks in astronomy which present an overview of the field at a
non-technical level. A classic one (though by now becoming of age)
and an excellent presentation of astronomy is
-
F. Shu: The Physical Universe: An Introduction to Astronomy , University Science Books, Sausalito, 1982.
Turning to more technical books, at about the
level of the present text, my favorite is
-
B.W. Carroll & D.A. Ostlie: An Introduction to Modern Astrophysics , Addison-Wesley, Reading, 2006;
its ∼ 1400 pages cover the whole range of
astronomy. The texts
-
M.L. Kutner: Astronomy: A physical perspective , Cambridge University Press, Cambridge, 2003,
-
J.O. Bennett, M.O. Donahue, N. Schneider & M. Voit: The Cosmic Perspective , Addison-Wesley, 2013,
also cover the whole field of astronomy. A text
with a particular focus on stellar and Galactic astronomy is
-
A. Unsöld & B. Baschek: The New Cosmos , Springer-Verlag, Berlin, 2002;
The book
-
M.H. Jones & R.J.A. Lambourne: An Introduction to Galaxies and Cosmology , Cambridge University Press, Cambridge, 2003
covers the topics described in this book and is
also highly recommended; it is less technical than the present
text.
D.2 More specific literature
More specific monographs and textbooks exist for
the individual topics covered in this book, some of which shall be
suggested below. Again, this is just a brief selection. The
technical level varies substantially among these books and, in
general, exceeds that of the present text.
Astrophysical
processes:
-
M. Harwit: Astrophysical Concepts , Springer, New-York, 2006,
-
G.B. Rybicki & A.P. Lightman: Radiative Processes in Astrophysics , John Wiley & Sons, New York, 1979,
-
F. Shu: The Physics of Astrophysics I: Radiation , University Science Books, Mill Valley, 1991,
-
F. Shu: The Physics of Astrophysics II: Gas Dynamics , University Science Books, Mill Valley, 1991,
-
S.N. Shore: The Tapestry of Modern Astrophysics , Wiley-VCH, Berlin, 2002,
-
D.E. Osterbrock & G.J. Ferland: Astrophysics of Gaseous Nebulae and Active Galactic Nuclei , University Science Books, Mill Valley, 2005.
Furthermore, there is a three-volume set of
books,
-
T. Padmanabhan: Theoretical Astrophysics: I. Astrophysical Processes. II. Stars and Stellar Systems. III. Galaxies and Cosmology , Cambridge University Press, Cambridge, 2000.
Galaxies and
gravitational lenses:
-
L.S. Sparke & J.S. Gallagher: Galaxies in the Universe: An Introduction , Cambridge University Press, Cambridge, 2007,
-
J. Binney & M. Merrifield: Galactic Astronomy , Princeton University Press, Princeton, 1998,
-
J. Binney & S. Tremaine: Galactic dynamics , Princeton University Press, Princeton, 2008,
-
R.C. Kennicutt, Jr., F. Schweizer & J.E. Barnes: Galaxies: Interactions and Induced Star Formation , Saas-Fee Advanced Course 26, Springer-Verlag, Berlin, 1998,
-
B.E.J. Pagel: Nucleosynthesis and Chemical Evolution of Galaxies , Cambridge University Press, Cambridge, 2009,
-
F. Combes, P. Boissé, A. Mazure & A. Blanchard: Galaxies and Cosmology , Springer-Verlag, 2004,
-
P. Schneider, J. Ehlers & E.E. Falco: Gravitational Lenses , Springer-Verlag, New York, 1992.
-
P. Schneider, C.S. Kochanek & J. Wambsganss: Gravitational Lensing: Strong, Weak & Micro , Saas-Fee Advanced Course 33, G. Meylan, P. Jetzer & P. North (Eds.), Springer-Verlag, Berlin, 2006.
Active
galaxies:
-
B.M. Peterson: An Introduction to Active Galactic Nuclei , Cambridge University Press, Cambridge, 1997,
-
R.D. Blandford, H. Netzer & L. Woltjer: Active Galactic Nuclei , Saas-Fee Advanced Course 20, Springer-Verlag, 1990,
-
J. Krolik: Active Galactic Nuclei , Princeton University Press, Princeton, 1999,
-
J. Frank, A. King & D. Raine: Accretion Power in Astrophysics , Cambridge University Press, Cambridge, 2012.
Cosmology:
-
M.S. Longair: Galaxy Formation , Springer-Verlag, Berlin, 2008,
-
J.A. Peacock: Cosmological Physics , Cambridge University Press, Cambridge, 1999,
-
T. Padmanabhan: Structure formation in the Universe , Cambridge University Press, Cambridge, 1993,
-
E.W. Kolb and M.S. Turner: The Early Universe , Addison Wesley, 1990,
-
S. Dodelson: Modern Cosmology , Academic Press, San Diego, 2003,
-
P.J.E. Peebles: Principles of Physical Cosmology , Princeton University Press, Princeton, 1993,
-
G. Börner: The Early Universe , Springer-Verlag, Berlin, 2003,
-
D.H. Lyth & A.R. Liddle: The Primordial Density Perturbation: Cosmology, Inflation and the Origin of Structure , Cambridge University Press, Cambridge, 2009.
D.3 Review articles, current literature, and journals
Besides textbooks and monographs, review articles
on specific topics are particularly useful for getting extended
information about a special field. A number of journals and series
exist in which excellent review articles are published. Among these
are Annual Reviews of Astronomy
and Astrophysics (ARA&A) and Astronomy & Astrophysics Reviews
(A&AR), both publishing astronomical articles only. In
Physics Reports (Phys.
Rep.) and Reviews of Modern
Physics (RMP), astronomical review articles are also
frequently found. Such articles are also published in the lecture
notes of international summer/winter schools and in the proceedings
of conferences; of particular note are the Lecture Notes of the
Saas-Fee Advanced Courses .
A very useful archive containing review articles on the topics
covered in this book is the Knowledgebase for Extragalactic
Astronomy and Cosmology, which can be found at
Original astronomical research articles are
published in the relevant scientific journals; most of the figures
presented in this book are taken from these journals. The most
important of them are Astronomy & Astrophysics (A&A), The
Astronomical Journal (AJ), The Astrophysical Journal (ApJ), Monthly
Notices of the Royal Astronomical Society (MNRAS), and Publications
of the Astronomical Society of the Pacific (PASP). Besides these, a
number of smaller, regional, or more specialized journals exist,
such as Astronomische Nachrichten (AN), Acta Astronomica (AcA), or
Publications of the Astronomical Society of Japan (PASJ). Some
astronomical articles are also published in the journals Nature and
Science. The Physical Review D and Physical Review Letters contain
an increasing number of papers on astrophysical cosmology.
Since many years now, the primary source of
astronomical information by far is the electronic archive
which is freely accessible. This archive, now
hosted at Cornell University and supported by the Simons Foundation
and the Allianz der deutschen Wissenschaftsorganisationen,
koordiniert durch TIB, MPG und HGF, has existed since 1992, with an
increasing number of articles being stored at this location. In
particular, in the fields of extragalactic astronomy and cosmology,
almost all articles that are published in the major journals can be
found in this archive. A large number of review articles and
coference proceedings are also available here.
The SAO/NASA Astrophysics Data System (ADS) is a
Digital Library portal for Astronomy and Physics, operated by the
Smithsonian Astrophysical Observatory (SAO) under a NASA grant. It
can be accessed via the Internet at, e.g.,
and it provides the best access to astronomical
literature. Besides tools to search for authors and keywords, ADS
offers also direct access to older articles that have been scanned.
The access to more recent articles, and to all articles in some
other journals, is restricted to IP addresses that are associated
with a subscription for the respective journals—but ADS also
contains a link to the article in the arXiv (if it has been posted
there), so also these articles are accessible.
E: Acronyms used
In this Appendix, we compile some of the acronyms
that are used, and references to the sections in which these
acronyms have been introduces or explained.
2dF(GRS)
|
–2 degree Field Galaxy Redshift Survey
(Sect. 8.1.2 )
|
2MASS
|
–Two Micron All Sky Survey (Sect.
1.4 )
|
AAS
|
–American Astronomical Society
|
AAT
|
–Anglo-Australian Telescope (Sect.
1.3.3 )
|
ACBAR
|
–Arcminute Cosmology Bolometer Array
|
–Receiver (Sect. 8.6.5 )
|
|
ACO
|
–Abell, Corwin & Olowin (catalogue of
clusters of galaxies, Sect. 6.2.1 )
|
ACS
|
–Advanced Camera for Surveys (HST
instrument—Sect. 1.3.3 )
|
ACT
|
–Atacama Cosmology Telescope (Sect.
8.6.6 )
|
ADAF
|
–Advection-Dominated Accretion Flow (Sect.
5.3.2 )
|
AGB
|
–Asymptotic Giant Branch (Sect.
3.5.2 )
|
AGN
|
–Active Galactic Nucleus (Sect. 5)
|
ALMA
|
–Atacama Large Millimeter/sub-millimeter
Array (Sect. 1.3.1 )
|
AMR
|
–Adaptive Mesh Refinement (Sect.
10.6.1 )
|
APEX
|
–Atacama Pathfinder Experiment (Sect.
1.3.1 )
|
ASP
|
–Astronomical Society of the Pacific
|
AU
|
–Astronomical Unit
|
BAL
|
–Broad Absorption Line (-Quasar, Sect.
5.7 )
|
BAOs
|
–Baryonic Acoustic Oscillations (Sect.
7.4.3 )
|
BATSE
|
–Burst And Transient Source Experiment
(CGRO instrument, Sect. 9.7 )
|
BBB
|
–Big Blue Bump (Sect. 5.4.1 )
|
BBN
|
–Big Bang Nucleosynthesis (Sect.
4.4.5 )
|
BCD
|
–Blue Compact Dwarf (Sect. 3.2.1 )
|
BCG
|
–Brightest Cluster Galaxy (Sect.
6.2.4 )
|
BH
|
–Black Hole (Sect. 5.3.5 )
|
BLR
|
–Broad Line Region (Sect. 5.4.2 )
|
BLRG
|
–Broad Line Radio Galaxy (Sect.
5.2.4 )
|
BOOMERANG
|
–Balloon Observations Of Millimetric
Extragalactic Radiation and Geophysics (Sect. 8.6.4 )
|
BTP diagram
|
–Baldwin–Phillips–Terlevich diagram (Sect.
5.4.3 )
|
CBI
|
–Cosmic Background Imager (Sect.
8.6.5 )
|
CCAT
|
–Cerro Chajnantor Atacama Telescope (Chap.
11 )
|
CCD
|
–Charge Coupled Device
|
CDF
|
–Chandra Deep Field (Sect. 9.2.1 )
|
CDM
|
–Cold Dark Matter (Sect. 7.4.1 )
|
CERN
|
–Conseil European pour la Recherché
Nucleaire
|
CfA
|
–Harvard-Smithsonian Center for
Astrophysics
|
CFHT
|
–Canada-France-Hawaii Telescope (Sect.
1.3.3 )
|
CFRS
|
–Canada-France Redshift Survey (Sect.
8.1.2 )
|
COSMOS
|
–Cosmological Evolution Survey (Sect.
9.2.1 )
|
CTIO
|
–Cerro Tololo Inter-American
Observatory
|
CXB
|
–Cosmic X-ray Background (Sect.
9.5.3 )
|
DASI
|
–Degree Angular Scale Interferometer
(Sect. 8.6.4 )
|
DES
|
–Dark Energy Survey (Chap. 11 )
|
DIRBE
|
–Diffuse Infrared Background Experiment
(instrument onboard COBE)
|
DLA system
|
–Damped Lyman Alpha system (Sect.
9.3.4 )
|
DRG
|
–Distant Red Galaxy (Sect 9.1.3)
|
dSph
|
–dwarf Spheroidal (Sect. 3.2.1 )
|
DSS
|
–Digital Sky Survey (Sect. 1.4 )
|
ECDFS
|
–Extended Chandra Deep Field South (Sect.
9.3.3 )
|
EdS
|
–Einstein–de Sitter (Sect. 4.3.4 )
|
E-ELT
|
–European Extremely Large Telescope (Chap.
11 )
|
EMSS
|
–Extended Medium Sensitivity Survey (Sect.
6.4 .5)
|
EPIC
|
–European Photon Imaging Camera (XMM-Newton
instrument)
|
ERO
|
–Extremely Red Object (Sect. 9.3.2 )
|
EROS
|
–Expérience pour la Recherche d’Objets
Sombres (microlensing collaboration, Sect. 2.5 )
|
ESA
|
–European Space Agency
|
ESO
|
–European Southern Observatory
|
FFT
|
–Fast Fourier Transform (Sect.
7.5.3 )
|
FIR
|
–Far Infrared
|
FIRAS
|
–Far Infrared Absolute Spectrophotometer
(instrument onboard COBE; see Fig. 4.3 )
|
FJ
|
–Faber–Jackson (Sect. 3.4.2 )
|
FOC
|
–Faint Object Camera (HST instrument)
|
FORS
|
–Focal Reducer / Low Dispersion
Spectrograph (VLT instrument)
|
FOS
|
–Faint Object Spectrograph (HST
instrument)
|
FP
|
–Fundamental Plane (Sect. 3.4.3 )
|
FR (I/II)
|
–Fanaroff–Riley Type (Sect. 5.1.2 )
|
FUSE
|
–Far Ultraviolet Spectroscopic Explorer
(Sect. 1.3.4 )
|
FWHM
|
–Full Width Half Maximum (Sect.
5.1.4 )
|
GALEX
|
–Galaxy Evolution Explorer (Sect.
1.3.4 )
|
GBM
|
–Gamma-ray Burst Monitor (instrument on
Fermi—Sect. 1.3.6 )
|
GC
|
–Galactic Center (Sects. 2.3 , 2.6)
|
GEMS
|
–Galaxy Evolution from Morphology and
Spectral Energy Distributions (Sect. 9.2.1 )
|
GGL
|
–Galaxy-Galaxy Lensing (Sect. 7.7 )
|
GMT
|
–Giant Magellan Telescope (Chap.
11 )
|
GOODS
|
–Great Observatories Origins Deep Survey
(Sect. 9.2.1 )
|
GR
|
–General Relativity
|
GRB
|
–Gamma-Ray Burst (Sect. 1.3.5 , 9.7)
|
GTC
|
–Gran Telescopio Canarias (Sect.
1.3.3 )
|
GUT
|
–Grand Unified Theory (Sect. 4.5.3 )
|
Gyr
|
–Gigayear = 10 9 years
|
GZK
|
–Greisen–Zatsepin–Kuzmin (Sect.
2.3.4 )
|
HB
|
–Horizontal Branch
|
HCG
|
–Hickson Compact Group (catalogue of galaxy
groups, Sect. 6.2.3 )
|
HDF(N/S)
|
–Hubble Deep Field (North/South) (Sect.
1.3.3 , 9.2.1)
|
HDM
|
–Hot Dark Matter (Sect. 7.4.1 )
|
HEAO
|
–High Energy Astrophysical Observatory
(Sect. 1.3.5 )
|
H.E.S.S.
|
–High Energy Stereoscopic System (Sect.
1.3.6 )
|
HFI
|
–High-Frequency Instrument (onboard the
Planck satellite; Sect. 8.6.6 )
|
HIFI
|
–Heterodyne Instrument for the Far Infrared
(Herschel instrument—Sect. 1.3.2 )
|
HLS
|
–Herschel Lensing Survey (Sect.
9.2.3 )
|
HRD
|
–Hertzsprung–Russell Diagram
(Appendix A.4.4 )
|
HRI
|
–High Resolution Imager (ROSAT
instrument)
|
HST
|
–Hubble Space Telescope (Sect.
1.3.3 )
|
HVC
|
–High Velocity Cloud (Sect. 2.3.6 )
|
HUDF
|
–Hubble Ultradeep Survey (Sect.
9.2.1 )
|
IAU
|
–International Astronomical Union
|
ICL
|
–IntraCluster Light (Sect. 6.3.4 )
|
ICM
|
–Intra-Cluster Medium (Chap. 6 )
|
IFU
|
–Integral Field Unit (Sect. 1.3.3 )
|
IGM
|
–Intergalactic Medium (Sect. 10.3 )
|
IMF
|
–Initial Mass Function (Sect. 3.5.1 )
|
IoA
|
–Institute of Astronomy (Cambridge)
|
IR
|
–Infrared (Sect. 1.3.2 )
|
IRAC
|
–Infrared Array Camera (instrument on
Spitzer—Sect. 1.3.2 )
|
IRAS
|
–Infrared Astronomical Observatory (Sect.
1.3.2 )
|
IRS
|
–Infrared Spectrograph (instrument on
Spitzer—Sect. 1.3.2 )
|
ISM
|
–Interstellar Medium
|
ISO
|
–Infrared Space Observatory (Sect.
1.3.2 )
|
ISW effect
|
–Integrated Sachs–Wolfe effect (Sect.
8.6.1 )
|
IUE
|
–International Ultraviolet Explorer (Sect.
1.3.4 )
|
IVC
|
–Intermediate-Velocity Cloud (Sect.
2.3.7 )
|
JCMT
|
–James Clerk Maxwell Telescope (Sect.
1.3.1 )
|
JVAS
|
–Jodrell Bank-VLA Astrometric Survey
(Sect. 3.11.3 )
|
JWST
|
–James Webb Space Telescope (Chap.
11 )
|
KAO
|
–Kuiper Airborne Observatory (Sect.
1.3.2 )
|
KiDS
|
–KiLO Degree Survey (Chap. 11 )
|
LAB
|
–Leiden-Argentine-Bonn (Sect. 1.4 )
|
LAE
|
–Lyman Alpha Emitter (Sect 9.1.3)
|
LAT
|
–Large Area Telescope (instrument on
Fermi—Sect. 1.3.6 )
|
LBG
|
–Lyman-Break Galaxy (Sect. 9.1.1 )
|
LBT
|
–Large Binocular Telescope (Sect.
1.3.3 )
|
LCRS
|
–Las Campanas Redshift Survey (Sect.
8.1.2 )
|
LFI
|
–Low-Frequency Instrument (onboard the
Planck satellite; Sect. 8.6.6 )
|
LHC
|
–Large Hadron Collider
|
LINER
|
–Low-Ionization Nuclear Emission-Line
Region (Sect. 5.2.3 )
|
LIRG
|
–Luminous InfraRed Galaxy (Sect.
9.4.1 )
|
LISA
|
–Laser Interferometer Space Antenna (Chap.
11 )
|
LMC
|
–Large Magellanic Cloud
|
LMT
|
–Large Millimeter Telescope (Chap.
11 )
|
LOFAR
|
–Low Frequency Array (Chap. 11 )
|
LSB galaxy
|
–Low Surface Brightness galaxy (Sect.
3.3.2 )
|
LSR
|
–Local Standard of Rest (Sect.
2.4.1 )
|
LSS
|
–Large-Scale Structure (Chap. 8 )
|
LSST
|
–Large Synoptic Survey Telescope (Chap.
11 )
|
MACHO
|
–Massive Compact Halo Object (and
collaboration of the same name, Sect. 2.5 )
|
MAGIC
|
–Major Atmospheric Gamma-ray Imaging
Cherenkov telescope (Sect. 1.3.6 )
|
MAMBO
|
–Max-Planck Millimeter Bolometer (Sect.
9.3.3 )
|
MAXIMA
|
–Millimeter Anisotropy Experiment Imaging
Array (Sect. 8.6.4 )
|
MDM
|
–Mixed Dark Matter (Sect. 7.4.2 )
|
MIPS
|
–Multiband Imaging Photometer for Spitzer
(instrument on Spitzer— Sect. 1.3.2 )
|
MIR
|
–Mid-Infrared
|
MLCS
|
–Multi-Color Light Curve Shape (Sect.
3.9.4 )
|
MMT
|
–Multi-Mirror Telescope
|
MOND
|
–Modified Newtonian Dynamics (Chap.
11 )
|
MS
|
–used for the ‘Main Sequence’ of stars, or
the ‘Millennium Simulation’ (Sect. 7.5.3 )
|
MW
|
–Milky Way
|
MXXL
|
–Millennium XXL simulation (Sect.
7.5.3 )
|
NAOJ
|
–National Astronomical Observatory of
Japan
|
NFW
|
–Navarro, Frenk & White (-profile,
Sect. 7.6.1 )
|
NGC
|
–New General Catalog (Chap. 3 )
|
NGP
|
–North Galactic Pole (Sect. 2.1 )
|
NICMOS
|
–Near Infrared Camera and Multi-Object
Spectrometer (HST instrument— Sect. 1.3.3 )
|
NIR
|
–Near Infrared
|
NLR
|
–Narrow Line Region (Sect. 5.4.3 )
|
NLRG
|
–Narrow Line Radio Galaxy (Sect.
5.2.4 )
|
NOAO
|
–National Optical Astronomy
Observatory
|
NRAO
|
–National Radio Astronomy Observatory
|
NTT
|
–New Technology Telescope (Sect.
1.3.3 )
|
NVSS
|
–NRAO VLA Sky Survey (Sect. 1.4 )
|
OGLE
|
–Optical Gravitational Lensing Experiment
(microlensing collaboration, Sect. 2.5 )
|
OVV
|
–Optically Violently Variable (Sect.
5.2.5 )
|
PACS
|
–Photodetector Array Camera and
Spectrometer (Herschel instrument— Sect. 1.3.2 )
|
PL
|
–Period-Luminosity (Sect. 2.2.7 )
|
PLANET
|
–Probing Lensing Anomalies Network
(microlensing collaboration, Sect. 2.5 )
|
PM
|
–Particle-Mesh (Sect. 7.5.3 )
|
P 3 M
|
–Particle-Particle Particle-Mesh (Sect.
7.5.3 )
|
PN
|
–Planetary Nebula
|
POSS
|
–Palomar Observatory Sky Survey (Sect.
1.4 )
|
PSF
|
–Point Spread Function
|
PSPC
|
–Position-Sensitive Proportional Counter
(ROSAT instrument)
|
QCD
|
–Quantum Chromodynamics (Sect.
4.4.1 )
|
QSO
|
–Quasi-Stellar Object (Sect. 5.2.1 )
|
RASS
|
–ROSAT All-Sky Survey (Sect. 6.4 .5)
|
RCS
|
–Red Cluster Sequence (Sect. 6.8 )
|
REFLEX
|
–ROSAT-ESO Flux-Limited X-Ray survey
|
RGB
|
–Red Giant Branch (Sect. 3.5.2 )
|
ROSAT
|
–Roentgen Satellite (Sect. 1.3.5 )
|
SAO
|
–Smithsonian Astrophysical
Observatory
|
SCUBA
|
–Sub-millimeter Common-User Bolometer Array
(Sect. 1.3.1 )
|
SDSS
|
–Sloan Digital Sky Survey (Sects.
1.4 , 8.1.2)
|
SFR
|
–Star Formation Rate (Sect. 9.6.1 )
|
SGP
|
–South Galactic Pole (Sect. 2.1 )
|
SIS
|
–Singular Isothermal Sphere (Sect.
3.11.2 )
|
SKA
|
–Square Kilometer Array (Chap.
11 )
|
SLACS
|
–Sloan Lens Advanced Camera for Surveys
(Sect. 3.11.3 )
|
SMBH
|
–Supermassive Black Hole (Sect.
5.3 )
|
SMC
|
–Small Magellanic Cloud
|
SMG
|
–Sub-Millimeter Galaxy (Sect. 9.3.3 )
|
SN(e)
|
–Supernova(e) (Sect. 2.3.2 )
|
SNR
|
–Supernova Remnant
|
SOFIA
|
–Stratospheric Observatory for Infrared
Astronomy (Sect. 1.3.2 )
|
SPH
|
–Smooth Particle Hydrodynamics (Sect.
10.6.1 )
|
SPIRE
|
–Spectral and Photometric Imaging REceiver
(Herschel instrument— Sect. 1.3.2 )
|
SPT
|
–South Pole Telescope (Sect. 1.3.1 )
|
SQLS
|
–SDSS Quasar Lens Search (Sect.
3.11.3 )
|
STIS
|
–Space Telescope Imaging Spectrograph (HST
instrument)
|
STScI
|
–Space Telescope Science Institute (Sect.
1.3.3 )
|
SZ
|
–Sunyaev–Zeldovich (-effect, Sect.
6.4 .4)
|
TDE
|
–Tidal Disruption Event (Sect.
5.5.6 )
|
TeVeS
|
–Tensor-Vector-Scalar (Chap. 11 )
|
TF
|
–Tully–Fisher (Sect. 3.4 )
|
TMT
|
–Thirty Meter Telescope (Chap.
11 )
|
TP-AGB star
|
–Thermally Pulsating AGB star (Sect.
3.5.5 )
|
UDF
|
–Ultra Deep Field (Sect. 9.2.1 )
|
UHECRs
|
–Ultra-High Energy Cosmic Rays (Sect.
2.3.4 )
|
ULIRG
|
–Ultraluminous Infrared Galaxy (Sect.
9.3.1 )
|
ULX
|
–Ultraluminous Compact X-ray Source (Sect.
9.3.1 )
|
UV
|
–Ultraviolet
|
VISTA
|
–Visible and Infrared Survey Telescope
(Sect. 1.3.3 )
|
VLA
|
–Very Large Array (Sect. 1.3.1 )
|
VLBA
|
–Very Long Baseline Array (Sect.
1.3.1 )
|
VLBI
|
–Very Long Baseline Interferometer (Sect.
1.3.1 )
|
VLT
|
–Very Large Telescope (Sect. 1.3.3 )
|
VST
|
–VLT Survey Telescope (Sect. 1.3.3 )
|
VVDS
|
–VIMOS VLT Deep Survey (Sect. 8.1.2 )
|
WD
|
–White Dwarf (Sect. 2.3.2 )
|
WDM
|
–Warm Dark Matter (Sect. 7.8 )
|
WFIRST
|
–Wide Field Infrared Space Telescope
(Chap. 11 )
|
WIMP
|
–Weakly Interacting Massive Particle
(Sect. 4.4.3 )
|
WISE
|
–Wide-field Infrared Survey Explorer(Sect.
1.3.2 )
|
WFC3
|
–Wide Field Camera 3 (HST instrument—Sect.
1.3.3 )
|
WFI
|
–Wide Field Imager (camera at the ESO/MPG
2.2m telescope, La Silla, Sect. 6.6.2 )
|
WFPC2
|
–Wide Field and Planetary Camera 2 (HST
instrument– Sect. 1.3.3 )
|
WMAP
|
–Wilkinson Microwave Anisotropy Probe
(Sect. 8.6.5 )
|
XDF
|
–eXtremely Deep Field (Sect. 9.2.1 )
|
XMM
|
–X-ray Multi-Mirror Mission (Sect.
1.3.5 )
|
XRB
|
–X-Ray Background (Sect. 9.5.3 )
|
ZAMS
|
–Zero Age Main Sequence (Sect.
3.5.2 )
|
F: Solutions to problems
Solution to
1.1. If the object at current distance D had a constant velocity v = H 0 D for all times, it needed a time
to reach separation D . This time is independent of
D . Using ( 1.7 ), we find that
where we used that 1 yr = 3. 16 × 10 7 s. For a value
of h ≈ 0. 71, this time is
comparable to, but slightly larger than the age of the oldest
stars. Light can propagate a distance c ∕ H 0 over the time-scale
H 0
−1 , where



Solution to
1.2. The number of galaxies in a sphere of radius
is
, where
.
Thus,
The number of these galaxies per square degree on the sky is
obtained by dividing N by
the solid angle of the sky, which is 4 π steradian. Since 180 ∘
corresponds to π rad, we
have that
, so that the full sky has a solid angle of
. This yields a number density of ∼ 2 × 10 3 deg
−2 .






To calculate the fraction of the sky covered by
the luminous region of these galaxies, we consider first a thin
spherical shell of radius r
and thickness d r around
us. In this shell, there are d N = 4 π r 2 n 0 d r galaxies, each of them subtending a
solid angle of π R
2 ∕ r
2 , where R
= 10 kpc is the radius of the luminous region. Thus, the solid
angle covered by all galaxies in the shell is d ω = 4 π 2 R 2 n 0 d r . The solid angle covered by all
galaxies within distance r
0 is obtained by integrating this expression over
r ,
The fraction of the sky covered by these galaxies is
.


Solution to
1.3.
(1)
The mean baryon density of the Universe is
. Making use of ( 1.14 ), this yields
. The estimate to the local mass density yields
. Thus, ρ ∗local
∕ ρ b ≈ 1. 6 ×
10 8 .



(2)
According to ( 1.1 ), the mass of the Galaxy inside
R 0 is
, yielding a mean density
within R 0 of
The mean matter density of the Universe is
. With Ω m
≈ 0. 3, this yields




(3)
The mean number of baryons N b in the box is the volume
of the box times the mean number density of baryons. The latter is
given by the mean mass density ρ b of baryons in the
Universe, divided by the mass per baryon, which is m b ≈ 1. 7 × 10
−24 g. Thus, making use of value of ρ b derived above,
Thus, the mean baryon density in the Universe is about
.


Solution
to 1.4.
(1)
Differentiating the ansatz for r ( t ), one finds
, and
.
Inserting this into the equation of motion yields
The powers of the time-dependent term must be the same one both
sides, yielding α − 2 = −2
α , or α = 2∕3. Equating the prefactors then
yields
The solution r (
t ) has infinite radius for
t → − ∞ , but the inflow velocity
tends to zero as t → − ∞ . At t = 0, r (0) = r 0 , but the velocity
is finite and determined by M and r 0 . For t = t f , the radius shrinks to
zero.







(2)
Replacing the mass by the mean initial density
of the sphere leads to
so that the time-scale t
f depends only on the initial density. We can now
compare t f to
the orbital time t
orb = 2 π R
0 ∕ V
0 . Using ( 1.1 ), and replacing M by the mean density, we get
.
Hence,
. Thus, t f
≈ 1. 75 × 10 7 yr.




(3)
Using the mean matter density of the Universe,
, as given by (
1.10 ), we obtain
Hence, t f is
very similar to the estimated age of the Universe, H 0 −1 , and
agrees with the age of the Einstein–de Sitter model ( 1.13 ), for which Ω m = 1. Indeed, the
expansion history of the Einstein–de Sitter model follows exactly
the same equation of motion, except that the relevant solution is
an expanding one, r (
t ) ∝ t 2∕3 , which is obtained
from our infalling solution by inverting the arrow of time (note
that the equation of motion is invariant against t → − t ), and shifting the origin of the
time axis.


Solution to
2.1. The angular diameter is δ = 3476∕385000 ≈ 9. 03 × 10
−3 . To convert this to degrees, we recall that
π = 180 ∘ , so
that
. The solid angle covered by the Moon is (0. 517 ∘ )
2 π
∕4 ≈ 0. 21 deg 2 , so the fraction it covers of the full
sky is (cf. problem 1.2) 0. 21∕41253 ≈ 5. 1 × 10 −6
.

Solution to
2.2. The total energy radiated throughout the galaxy’s
lifetime is E =
L t , where t = 10 10 yr is the
assumed age. The energy generated is the mass that is converted
into helium in nuclear fusion, times the energy released per unit
mass. The former is Y
M , where Y is the helium mass fraction generated
by nuclear fusion. The energy released per unit mass is
ε c 2 , where ε is the efficiency of this energy
generating process, given by the ratio of the binding energy per
nucleon in helium ( ∼ 28 MeV∕4 = 7 MeV) and the mass per nucleon,
m nuc ≈
m p ≈ 938 MeV;
i.e., ε ≈ 0. 008. Thus,
the total energy released is E = M Y ε c 2 . Equating this to
L t , we obtain
where we used the mass-to-light ratio. Inserting the values for
M ⊙ ≈ 2 × 10
33 g and L
⊙ ≈ 3. 8 × 10 33 erg∕s, and using
t ≈ 3. 1 × 10
17 s, we obtain Y ≈ 2. 7 % . This value is lower by a factor of
ten than the observed helium abundance. On the other hand, adding
this value to the helium abundance from BBN yields a result which
is very close to the currently observed helium abundance in local
galaxies.

Solution to
2.3. According to the Kepler rotation,
, a constant V implies M ( R ) ∝ R . The relation between M and ρ is
Differentiating this w.r.t. R yields
. On
the other hand, M (
R ) ∝ R implies
, so that
ρ ( r ) r 2 = const., or
ρ ( r ) ∝ r −2 .




Solution to
2.4. Light at the Solar limb is deflected by
,
which is far smaller than the angular radius θ ⊙ = 16 ′ of the Sun. Hence, if we
consider a cone of light rays with vertex at the Earth and an
opening angle of θ
⊙ , this cone will continue to diverge after being
deflected at the Solar limb. If the Sun had a larger distance
D , its angular radius
would be smaller, namely
. If θ equals
, the rays of the cone with
opening angle θ at Earth
would be parallel after light deflection at the Solar limb, and if
the distance was slightly larger they would converge after
deflection and go through a common focus. If a source was placed at
this focus, the Sun would then produce an Einstein ring. Displacing
the source slightly, the ring breaks up into a pair of images, with
angular separation 2 θ .
Thus, the minimum distance for lensing is
, and the image splitting at the minimum distance would be
.





Solution to
2.5. Kepler rotation yields
, which we
rewrite as
With M • ≈ 4 ×
10 6 M
⊙ , we obtain at r = 4 pc a rotational velocity of
.
Hence, at about this radius, the Keplerian rotation velocity around
the black hole equals the velocity dispersion of the stellar
cluster.



Solution to
2.6. A light ray from the GC to us which is scattered in the
screen at radius R has a
geometric length of
, where L 1 is the length of the ray
path from us to the point R
in the screen, and L
2 the length of the path from the GC to that point.
Trigonometry then yields
, where we made use of the fact that R ∕ D ≪ 1 and used a first-order Taylor
expansion of the square root. Similarly,
. Thus, the total length of the ray path is
The light-travel time along this ray is L ∕ c . If we define t as the excess of this light-travel
time relative to a straight ray, we obtain
where in the final step we used D ≪ R 0 , or D sc ≈ R 0 . Hence, R 2 ≈ 2 cDt . Differentiating both sides w.r.t.
t , we get
, or
. This apparent velocity is larger
than the speed of light, since R ≪ D . If the scattering screen is located
behind the Galactic center, the situation is very similar to the
one described here.

![$$L_{2} = \sqrt{D^{2 } + R^{2}} = D\sqrt{1 + (R/D)^{2}} \approx D\left [1 + R^{2}/(2D^{2})\right ]$$](A129044_2_En_BookBackmatter_OnlinePDF_IEq70.gif)





Solution to
3.1.
(1)
We consider the central square arcsecond of the
galaxy as a source; its apparent magnitude (all magnitudes
considered here are in the B-band) is m = μ 0 arcsec 2
= 21. 5. Thus, its absolute magnitude is
. The corresponding
luminosity of that source is
. The central square arcsecond corresponds to an area of (
D 1 ′ ′ ) 2 , and so the
central surface brightness is
. Thus,




(2)
At the distance of 16 Mpc, 50 ′ ′ correspond to h R ≈ 4 kpc. The luminosity is
obtained by integrating the surface brightness over the disk,
and by inserting numbers,


Solution to
3.2. According to the assumptions, the mass function of the
stars is n ( m ) d m ∝ m − α d m , with α = 2. 35. If half the mass is
contained in stars with mass m L ≤ m ≤ m m 50 , then
so that the result for m
m 50 is
independent of the normalization of the mass function. Integration
then yields
Solving for m
m 50 yields
or m m 50 ≈ 0. 55 M ⊙ .



To calculate m L 50 , we need to satisfy
The same steps than lead to the result
or m L 50 ≈ 46 M ⊙ . We thus see that a
stellar population with such a mass spectrum contains most of its
mass in the low-mass stars, whereas most of the luminosity is due
to the highest-mass stars.


Solution to
3.3. The volume contained in a solid angle ω out to distance D is
. If
is the supernova event rate per unit volume,
the observed SNe rate in the solid angle ω up to distance D is
. Inserting numbers,
. Since
, we can write
,
so that
. Thus, one needs to survey ∼ 80 deg 2 to find 10 nearby
SNe per year.







Solution to
3.4.
(1)
The volume of a cone with height D lim and opening solid
angle ω is

(2)
If the luminosity of a source is larger than
L = 4 π S min D lim 2 , its
flux is larger than S
min even out to distances D lim , and so they can be
seen throughout the search volume. On the other hand, if their
luminosity is smaller than this value, they can only be seen out to
a distance
. Thus,


(3)
If Φ (
L ) d L is the number density of galaxies
with luminosity within d L
of L , then we observe
galaxies in the survey with luminosity ≤ L . Differentiation of this relation
then immediately yields the desired result.

Solution to
3.5.
(1)
(2)
The corresponding lens equation reads
. Dividing the lens equation by λ , we obtain
. That means that the new mass distribution κ λ yields the same image positions
as the original mass model, provided the source position is changed
from
to
. Since the source
position in unobservable, this shift in the source plane can not be
observed.
![$${\boldsymbol \beta }={\boldsymbol \theta } -{\boldsymbol \alpha }_{\lambda }({\boldsymbol \theta }) =\lambda \left [{\boldsymbol \theta }-{\boldsymbol \alpha }({\boldsymbol \theta })\right ]$$](A129044_2_En_BookBackmatter_OnlinePDF_IEq87.gif)



(3)
The magnification is given by ( 3.68 ), and thus reads for the
modified mass model κ
λ
Hence, the magnification of all images is changed with this new
mass model; however, the magnification ratios between images stays
the same, and thus the predicted flux ratios.

Solution to
4.1. At BBN, T
≈ 0. 1 MeV ≈ 10 9 K ≈ 4 × 10 8 T 0 , where T 0 is the current
temperature of the CMB. Hence, BBN happens at a scale factor of
a BBN ≈ 2. 5 ×
10 −9 . The current baryon density is determined by (
4.68 ) and the critical density,
ρ b (0) ≈ 0. 02
h −2
ρ cr , and the
baryon density at BBN is then
. Using ( 4.15 ), we find
. This density is many orders of magnitude lower than in the center
of the Sun, or in other stars.


Nuclear burning in stars is slow because the
thermal energy of protons (i.e., temperature) is too low to allow
overcoming the Coulomb barrier, that is the electrostatic repulsion
between equally charged particles. Only very rarely does this
process happen; quantum-mechanically, it occurs by a process called
‘tunneling’. In contrast, the temperature during BBN was high,
which makes it far easier to beat electrostatic repulsion. Second,
and most important, stars contain essentially no free neutrons, in
contrast to the situation at BBN; the formation of deuterium at BBN
thus did not involve a Coulomb barrier, only the transformation of
deuterium into helium.
The energy density ρ E released during BBN is given by
the mass fraction of nucleons that ended up in helium, which is
Y ≈ 0. 25, times the
binding energy per nucleon in helium, corresponding to ∼ 7 MeV,
times the number density of nucleon n nucl . The latter is given
by the baryon density at BBN, ρ b (BBN), divided by the
mass of the nucleon, which is about 940 MeV∕ c 2 . Thus,
The energy density of photons at BBN is given by
. Hence,
where we made use of the results from the first part of this
problem and used ( 4.27 ). Hence, the energy released
during BBN is totally negligible compared to the energy of the
photon gas.



Solution to
4.2. From the general definition ( 4.34 ), we need to calculate
We have (at the current epoch)
, and
, where the sum extends over all N species. Hence, with the definition
of ρ cr , we
obtain
. Using ( 4.34 ), this yields
.
With a pressureless matter component ( w = 0) and the vacuum energy with
, ( 4.35 ) is recovered.






Solution to
4.3. Since
, an expansion can
change into a contraction (or the opposite) only when H ( a ) = 0. For Ω Λ = 0, we have
. This expression is always positive for 0 < a < 1. From the first form of
E 2 , we see
that this expression is also positive for a > 1 if Ω m ≤ 1. If Ω m > 1, E 2 = 0 at
.



Including a finite Ω Λ , we have
. We rewrite this expression as
and see that all terms are non-negative for all a > 1 if Ω m ≤ 1, i.e., the universe
expands forever in the future. Using the form
, we see that all terms are positive provided 0 ≤ Ω Λ < 1.



Assume that at t ex , an expansion turns
into a contraction. Since
, we then have that
for t < t ex , and
for t > t ex . By integrating this
expression, we obtain for t
> 0:
and for t < 0
Combining these two equation yields
which implies
.







Solution to
4.4. The Friedmann equation in a flat universe reads
, with Ω Λ = 1 − Ω m . With the ansatz
a = v β , we find
, in terms of which
the expansion equation becomes
. The desired form is achieved by setting
, yielding
.




![$$\dot{v}^{2} = (9H_{0}^{2}\varOmega _{\mathrm{m}}/4)\left [1 + (\varOmega _{\varLambda }/\varOmega _{\mathrm{m}})v^{2}\right ]$$](A129044_2_En_BookBackmatter_OnlinePDF_IEq113.gif)
With the ansatz
,
, and we obtain
. In order for this equation to be valid, the time-dependence on
the r.h.s. must be of the form cosh 2 ( t ∕ t a ); this can be achieved
by choosing v 0
such that
, or
. The prefactors on both sides also must agree, which then
determines
.



![$$\left.(\varOmega _{\varLambda }/\varOmega _{\mathrm{m}})v_{0}^{2}\sinh ^{2}(t/t_{\mathrm{a}})\right ]$$](A129044_2_En_BookBackmatter_OnlinePDF_IEq117.gif)



Hence, the final solution reads
Considering the case t ≪
t a and making
use of sinh( x ) ≈
x for x ≪ 1, we find
, which agrees with ( 4.70 ). Hence, for times
t ≪ t a , the expansion law does
not contain Ω
Λ explicitly. On
the other hand, for t ≫
t a , using
sinh( x ) ≈ e
x ∕2 for
x ≫ 1, we obtain
. Hence, for late times, the universe expands exponentially. Note
that this last solution satisfies
,
the Friedmann equation for a Λ -dominated universe.




Finally, we consider the sign of the second
derivative of a . With
, we find in turn
,
, and thus
. The prefactor is always positive, and the term in brackets is
negative for t ≪
t a , since
cosh( x ) ≈ 1 for
x ≪ 1, and positive for
t ≫ t a , when both
. Hence, the solution describes the transition from an decelerating
universe to an accelerating one.



![$$\ddot{a} = 2v_{0}^{2}/(9v^{4/3}t_{\mathrm{a}}^{2})\left [3\sinh ^{2}(t/t_{\mathrm{a}}) -\cosh (t/t_{\mathrm{a}})\right ]$$](A129044_2_En_BookBackmatter_OnlinePDF_IEq127.gif)

Solution to
4.5. In this case, the Friedmann equation reads
. Using the same ansatz as in problem 4.4, we now need to choose
,
, and
, to obtain
For t ≪ t a , this becomes
, which has the t
1∕2 -dependence that we derived for the
radiation-dominated era. For t ≫ t a , the expansion is
exponential.






Solution to
4.6. From ( 4.39 ), we find
, where we used the relation between physical length and comoving
length, d r =
a d x . Hence,
Integrating both sides from some point along the ray, characterized
by the scale factor a or,
equivalently, the redshift
, and corresponding to the comoving
distance x ( z ), and using ( 4.33 ), one arrives at (
4.53 ).



Consider two light rays separated by a small
angle θ at the observer.
For a flat universe, the comoving separation between these two rays
is then given by L (
a ) = θ x . The physical separation is then
aL ( a ). According to the definition of
D A , we then
have
, which reproduces
( 4.54 ) for the case K = 0.

Solution to
4.7.
(1)
Consider the case K > 0 first. We note from (
4.85 ) that the parameter
θ corresponding to
t = t 1 is θ 1 = 0, since (
θ − sin θ ) is a monotonically increasing
function. The first of ( 4.85 ) then yields f ( t 1 ) = 0, i.e., the initial
condition f = 0 at
t = t 1 is satisfied by (
4.85 ). Denoting derivatives w.r.t.
t by a dot, those w.r.t.
θ by a prime, (
4.84 ) reads
. Differentiation yields
,
, so that
. The l.h.s. of ( 4.84 ) then becomes
The r.h.s. of ( 4.84 ) is
, which is seen to
agree with the above expression. Hence, ( 4.85 ) indeed solves ( 4.84 ) with the correct initial
condition. The case K
< 0 can be treated in the same way. For K = 0, ( 4.87 ) yields f ( t 1 ) = 0, as required, and
. Simple
algebra then shows that
. Considering the case
K > 0 again,
f attains a maximum where
cos θ has its minimum,
which occurs for θ =
π ; hence,
, at time
. Furthermore, for θ = 2
π , f = 0, which happens at time
.











(2)
For
, ( 4.33 ) reads
, which is
seen to have the same form as ( 4.84 ), with C = H 0 2
Ω m and
. Setting
t 1 = 0 then
yields a (0) = 0, the
correct initial condition for Friedmann expansion. Hence, with
these parameter values, ( 4.85 ) describes the expansion for
Ω m > 1, (
4.86 ) the expansion for
Ω m < 1, and
( 4.87 ) yields the EdS case,
. For Ω m > 1, the previous
results then show that
, occurring at time
, and collapse happens at t
coll = 2 t
max .


![$$\left.(1 -\varOmega _{\mathrm{m}})\right ]$$](A129044_2_En_BookBackmatter_OnlinePDF_IEq149.gif)




(3)
Differentiation of ( 4.84 ) w.r.t. t yields
, or
. The equation for the radius
of the sphere is
, so that we can identify
f with r , and C = 2 GM . The constant K is proportional to the (negative of
the) total energy of sphere,
, as the sum of specific
kinetic and potential energy. With r 0 = r (0), we can write
. The solution found
in problem 1.4 was
, which yields
, when using the value for t f derived in
problem 1.4. This solution corresponds to one where the sphere
at t = 0 has an initial
infall speed, such that the total energy of the sphere is zero, in
full analogy to a time-reversed Einstein–de Sitter model. Setting
the initial velocity to zero,
, yields
. In the context of the equation
of motion discussed above, the free-fall time is the time between
the maximum expansion and the time of collapse,
, which yields
. Replacing M by the mean
density of the sphere, we arrive at ( 4.88 ).






![$$K = (1 - t/t_{\mathrm{f}})^{-2/3}\left [2GM/r_{0} - 4r_{0}^{2}/(9t_{\mathrm{f}}^{2})\right ] = 0$$](A129044_2_En_BookBackmatter_OnlinePDF_IEq160.gif)




Solution to
4.8.
(1)
Conservation of kinetic plus potential energy
(per unit mass) yields
,
or
. At initial
time t 0 ,
r ( t 0 ) = r E ,
, so that
, which is
assumed to be positive in the following.




(2)
The equation of motion has the form (
4.84 ), with C = 2 GM E , hence the solution (
4.85 ) applies. Without loss of
generality, we set t
1 = 0, and denote the time when the object leaves the
Earth surface by t
0 . The corresponding parameter value θ 0 is found from the first
of ( 4.85 ),
, where we used the leading order of the Taylor expansion and
assumed that θ 0
≪ 1, which needs to be verified. Writing the initial velocity as a
fraction μ of the escape
velocity, v 0
= μ v esc , we
see that
. Hence, θ 0
≪ 1 if the initial velocity is sufficiently close to the escape
velocity, 1 − μ ≪ 1, which
will be assumed in the following (corresponding to the assumption
that the flight is ‘long’). Then using the second of ( 4.85 ), we obtain in the same manner
. Since
, we finally obtain
.





(3)
The return time t ret is given by
t coll − 2
t 0 , with
. Provided 1 − μ ≪ 1,
t coll ≫
t 0 , and so we
can approximate t
ret ≈ t
coll . Inserting numbers, we obtain
. Setting t ret
= 1 d, we find μ ≈ 0. 93,
whereas for t
ret = 1 yr, μ
≈ 0. 994. Hence, if one wants to have a really long flight, the
initial velocity must be extremely well tuned.


Solution to
4.9. The momentum behaves like
, where m is the rest mass of the
(non-relativistic) particle, and v its velocity as measured by a
comoving observer. The temperature T b is related to the mean
velocity dispersion as
. Thus,
T b ∝
v 2 ∝ (1 +
z ) 2 .


Solution to
4.10. Consider a cosmic epoch, characterized by the scale
factor a , at which the
neutrinos were ultra-relativistic; their momentum then was
. The mean energy of the thermal distribution
is about 3 T
ν , where
T ν ∼ 0. 7 T γ , according to ( 4.62 ). With
, we then get
. This
product is conserved, as shown by ( 4.47 ). When the temperature of the
universe drops below the neutrino rest mass, the momentum is
p = m ν v . Thus, we obtain for the
characteristic velocity of cosmic neutrinos




Solution to
4.11.
(1)
From ( 4.56 ), we see that
. The look-back time is
. The redshift where
is then determined by
, or
.

![$$\tau (z) = t_{0} - t(z) = \left [1 - (1 + z)^{-3/2}\right ]t_{0}$$](A129044_2_En_BookBackmatter_OnlinePDF_IEq182.gif)



(2)
The volume of the spherical shell is the product
of the surface of the sphere at redshift z and the physical thickness
c d t of the shell, corresponding to the
redshift interval d z . By
definition of the angular diameter distance, the surface is 4
π D A
2 ( z ), so that
Furthermore, from
and
, we find that
. Using (
4.57 ) and
, valid for the EdS model, one
finally finds



![$$\mathrm{d}t = -\mathrm{d}z/[(1 + z)H]$$](A129044_2_En_BookBackmatter_OnlinePDF_IEq188.gif)


(3)
The number of objects is
, where the comoving volume element d V com is related to the
physical (proper) volume element by
. Thus
Specializing the last result to z ≪ 1, using
, we then obtain
,
which is the number of objects in a sphere of radius c z ∕ H 0 .





Solution to
4.12. The reason for the absence of H 0 in ( 4.61 ) is that in the Friedmann
equation ( 4.18 ) curvature and cosmological
constant can be neglected at early times, and thus
. In other words,
, where the constant
C depends on the number of
relativistic species only—see ( 4.60 ). The first law of
thermodynamics ( 4.17 ) then yields that T ∝ 1∕ a [see ( 4.24 )], which leads to
, without any
reference to the current universe – neither to its expansion rate,
nor to its current temperature. The foregoing equation can be
solved with the ansatz T (
t ) = xt α , which yields
, and thus
,
.






The helium abundance depends on the density of
baryons (given that the density of photons as a function of
temperature is known). But
. Hence, the combination Ω
b h 2
determines the physical baryon density.

Solution to
4.13. The scattering optical depth is given by the
line-of-sight integral over the product of the Thompson scattering
cross section σ
T and the number density n e of free electrons,
From problem 4.11 we know that
(in this problem we
set 1 + z ≈ z , since all the contributions to the
integral comes from z
≳ 800). Since recombination happens in the matter-dominated epoch,
we have
. The number density of free electrons is equal to the number
density of protons n
p and the ionization fraction x . For n p , we have n p ∝ Ω b ρ cr z 3 ∝ Ω b h 2 z 3 , so that
Using ( 4.72 ), we see that the Ω ’s and h ’s drop out, and we obtain
which yields the correct functional dependence of ( 4.73 ).





Solution to
5.1.
(1)
(2)
The distribution of relativistic electrons is
, since E = γ m e c 2 , and a is a constant ∝ A . The synchrotron emissivity of this
distribution is obtained by integrating the emissivity of a single
electron over the electron distribution,
where we defined
. Changing the integration variable to
, with
then yields
Since the final integral is frequency-independent, we find that the
emitted spectrum is a power law with index
.







Solution to
5.2.
(1)
(2)
Consider first the case that all photons have the
same energy E
γ . The energy
loss of a relativistic electron scattering a photon then is on
average
, and the
time between scatterings is
. With U γ = n γ E γ and
,
the desired expression is obtained. Since this expression no longer
refers to the photon energy (or frequency), but only to the energy
density of photons, the same result is obtained for a spectral
distribution of photons.



Solution to
5.3. We write the temperature profile ( 5.13 ) as
. The specific intensity
is that of a blackbody with temperature T ( r ), so that
The emitted luminosity L
ν is then the
integral of the specific intensity over the surface of the disk,
where we neglected boundary effects by setting the integration
limits to 0 and ∞ . Writing
the exponent as
where in the last step we defined ν 0 , we see that
. Therefore,
, and

![$$\displaystyle{I_{\nu }(r) \propto \nu ^{3}\left [\exp \left ( \frac{h_{\mathrm{P}}\nu } {k_{\mathrm{B}}T(r)}\right ) - 1\right ]^{-1}\;.}$$](A129044_2_En_BookBackmatter_OnlinePDF_Equar.gif)





Solution to
5.4. At an accretion rate of
, the growth rate of the black hole is
, since
the fraction ε of the
accretion rate is converted into luminosity, and thus does not
contribute to the increase of the black hole mass. With (
5.15 ), one then finds
and the relation ( 5.47 ) is the solution of this
equation.



Inserting the specific values L = L edd , ε = 0. 1, and t = 10 9 yr, we find
M • (
t ) = M • (0) e 18
≈ 6. 6 × 10 7 M • (0), or M • ( t ) ≈ 6. 6 × 10 8
M ⊙ .
Solution to
5.5.
(1)
We assume all clouds to be at the characteristic
distance r from the SMBH;
each cloud covers a solid angle of π r c 2 ∕
r 2 . The
covering fraction is obtained by summing the solid angle over all
N c clouds and
dividing by 4 π , to find
.

(2)
The volume filling factor f V is the ratio of the total
volume of the clouds, N
c (4 π ∕3)
r c 3
to that of the BLR, (4 π
∕3) r 3 ,
.

(3)
Using
, one obtains
.
From f cov
= 0. 1 we get
.
Combining these two expressions, we obtain N c = 6. 4 × 10
10 and
. The total gas mass of the clouds is the product of the clouds’
volume times the electron density (that then yields the total
number of electrons in the clouds) times the average mass per
electron, which is about the proton mass (since there are about as
many electrons as nucleons). Hence, M c = n e V c m p . The volume is
, so that
. Hence, the total gas mass in the BLR is very small indeed.






Solution to
5.6.
(1)
According to the assumption, the optical light
from the AGN is
where we used the Solar luminosity
.
The optical light from the host galaxy is
. Using
, we
then find





(2)
With an Eddington ratio of 0. 1 and a typical
mass-to-light ratio of 3 in Solar units, the light from the AGN is
comparable to that of the host galaxy if the spheroidal fraction is
close to unity. For late-type galaxies, where f sph is considerably
smaller, the host galaxy will typically dominate the optical
emission—this is one of the reasons why a complete census of AGNs
in the optical is difficult to achieve. The AGN can outshine the
host galaxy only for apparently large Eddington ratios (e.g., when
beaming is involved), or if the SMBH is considerably larger than
the assumed scaling.
Solution to
5.7. If the tidal disruption distance
is smaller than the Schwarzschild radius of the SMBH, the star will
be swallowed by the black hole before it can be disrupted. Hence we
require R t
> r S .
This yields
. Dividing the expression by the Solar mass then yields
where we used the Solar radius R ⊙ ≈ 7 × 10 5
km and the Schwarzschild radius of the Sun, which is
2. 95 km.



Solution to
6.1. With
and Kepler’s law
, we find
. Since
, we see that
Comparing this to the free-fall time yields
.
Inserting
then yields
. Since the
inverse of the expansion rate H ( z ) is, up to factor of order unity,
the age of the Universe at that epoch, we see that t orbit ∼ t ( z ).


![$$t_{\mathrm{orbit}} = 2\pi [GM/r^{3}]^{-1/2}$$](A129044_2_En_BookBackmatter_OnlinePDF_IEq231.gif)


(E.1)



Solution to
6.2. The observed bolometric flux is, according to the
definition ( 4.50 ),
, and the angular
radius is
. Hence, the surface
brightness behaves like
. For the specific surface brightness I ν , the redshift-dependence can
be different, depending on the spectral properties of the source,
according to the necessary K-correction (see Sect. 5.6.1 ). With
and
S ν from ( 5.42 ), we obtain
, yielding
.






Solution to
7.1. If ρ is assumed
to be spatially constant, the continuity equation becomes
. Since the first term is independent of
, so must be the second, which
immediately implies that
can depend only on time.
Inserting
then yields
, or
. By insertion, we see
that the solution with ρ (
t 0 ) =
ρ 0 is given by
( 4.11 ).






Since
, a
solution of the Poisson equation is
so that
. For the terms on the l.h.s. of the Euler equation, we find
Combining these terms, we see that the Euler equation yields (
4.19 ), with P = 0.


![$$\nabla \varPhi = \left [(4\pi /3)G\rho -\varLambda /3\right ]{\boldsymbol r}$$](A129044_2_En_BookBackmatter_OnlinePDF_IEq249.gif)

Solution to
7.2.
(1)
(2)
We show next that
, with

![$$\displaystyle{I(a) =\int _{ 0}^{a} \frac{\mathrm{d}a^{{\prime}}} {[a^{{\prime}}\,H(a^{{\prime}})]^{3}}\;,}$$](A129044_2_En_BookBackmatter_OnlinePDF_Equba.gif)
is a solution of ( 7.15 ). We first note that
so that
Furthermore,
Collecting terms,
where the final equality was shown in the first part of this
problem.




(3)
For the EdS model, we have
,
and
. Inserting these expression as
D = H into ( 7.15 ), using
, shows that it is satisfied,
i.e., H is a solution of
the growth equation. Furthermore, specializing ( 7.17 ) to the EdS parameters yields
Inserting
into ( 7.15 ) shows that it indeed solves
the growth equation.






Solution to
7.3.
(1)
The relation between virial mass and virial
radius is given by ( 7.56 ), so that
Using the Hubble function for a flat universe, and transforming the
product GM into the
Schwarzschild radius yields
Inserting the values for the Schwarzschild radius of the Sun and
c ∕ H 0 , we find
With the estimates 3 × 81∕200 ≈ 1. 2, the number in the last
parenthesis yields approximately 1. 2 × 10 59 = 120 ×
10 57 . Since 5 3 = 125, the third root of
120 is very close to 5, so that
Thus, at z = 0, we obtain
and
for the galaxy and
cluster mass halo, respectively.

![$$\displaystyle{r_{200} = \left (\frac{2GM_{\odot }} {c^{2}} \, \frac{M} {M_{\odot }} \frac{c^{2}} {200\,H_{0}^{2}\left [\varOmega _{\mathrm{m}}(1 + z)^{3} +\varOmega _{\varLambda }\right ]}\right )^{1/3}.}$$](A129044_2_En_BookBackmatter_OnlinePDF_Equbf.gif)
![$$\displaystyle\begin{array}{rcl} & & r_{200} \approx \left ( \frac{M} {M_{\odot }}\right )^{1/3}\left [\varOmega _{\mathrm{ m}}(1 + z)^{3} +\varOmega _{\varLambda }\right ]^{-1/3}h^{-2/3} {}\\ & & \qquad \times \left (\frac{3 \times 10^{5}\;\;81 \times 10^{54}} {200} \right )^{1/3}\mathrm{cm}\;. {}\\ \end{array}$$](A129044_2_En_BookBackmatter_OnlinePDF_Equ55.gif)
![$$\displaystyle\begin{array}{rcl} r_{200}\!& \approx &\!5\times 10^{19}\mathrm{cm}\!\left ( \frac{M} {M_{\odot }}\right )^{1/3}\!\left [\varOmega _{\mathrm{ m}}(1+z)^{3}+\varOmega _{\varLambda }\right ]^{-1/3}\!h^{-2/3} \\ & \approx & 16\mathrm{pc}\left ( \frac{M} {M_{\odot }}\right )^{1/3}\left [\varOmega _{\mathrm{ m}}(1+z)^{3}+\varOmega _{\varLambda }\right ]^{-1/3}h^{-2/3}. {}\end{array}$$](A129044_2_En_BookBackmatter_OnlinePDF_Equ56.gif)
(E.2)


At z
= 2, we see that the matter term in the bracket dominates, since
, so that
, since 2 3 = 8. Thus, the virial radius at
z = 2 is about half the
size of that today, for fixed mass. Hence, halos of a given mass
are smaller at higher redshift, as expected, since the definition
of a halo—‘mean density equals 200 times critical density’—together
with the increase of the critical density with redshift implies
that higher-redshift halos have a larger mean density, i.e., they
are more compact.

![$$\left [\varOmega _{\mathrm{m}}(1 + z)^{3} +\varOmega _{\varLambda }\right ]^{-1/3} \approx 1/2$$](A129044_2_En_BookBackmatter_OnlinePDF_IEq262.gif)
For V
200 , we start from ( 7.58 ) and use the preceding result:
Hence, our two halos have virial velocities of 160 km∕s and
1600 km∕s, respectively, at redshift zero, independent of
h . At z = 2, the Ω -dependent term is
(recall the earlier estimate), hence
the corresponding virial velocities are 225 km∕s and 2250 km∕s,
respectively. That they are higher by a factor
is already clear from Kepler’s law, since
the radius is smaller by a factor 2, for fixed mass.
![$$\displaystyle\begin{array}{rcl} V _{200}& =& 10\,H(z)\,r_{200} \\ & =& 10\,h\,100\,\mathrm{km\,s^{-1}\,Mpc^{-1}}\sqrt{\varOmega _{\mathrm{ m}}(1+z)^{3}+\varOmega _{\varLambda }} \\ & \times & 16\mathrm{pc}\left ( \frac{M} {M_{\odot }}\right )^{1/3}\!\left [\varOmega _{\mathrm{ m}}(1+z)^{3}+\varOmega _{\varLambda }\right ]^{-1/3}h^{-2/3} \\ & =& 16\times 10^{-3}\mathrm{\frac{km} {s} }\left ( \frac{M} {M_{\odot }}\right )^{1/3}\!\left [\varOmega _{\mathrm{ m}}(1+z)^{3}+\varOmega _{\varLambda }\right ]^{1/6}h^{1/3}. {}\end{array}$$](A129044_2_En_BookBackmatter_OnlinePDF_Equ57.gif)
(E.3)


(2)
The mass M contained in a proper volume
V prop at
redshift z in a homogeneous
Universe is
. Correspondingly, the mass in the comoving volume
is
,
independent of redshift. Writing
and using the
definition of the virial radius, specialized to the current epoch,
we obtain
where r 200 is
the virial radius of the halo of mass M today. Thus,
Thus, the mass of our two halos were assembled from a volume
corresponding to a sphere of comoving radius of 1. 4 h −1 Mpc and 14
h −1 Mpc,
respectively.






(3)
According to ( 7.38 ), the scale factor
a at which a perturbation
of length scale L enters
the horizon is given as
where we made a first-order Taylor expansion, assuming a ≪ a eq , and used the
definition of a
eq . Note that we have just rederived ( 4.76 ). Using ( 4.28 ), this becomes
Thus, the perturbation that eventually led to the formation of our
galaxy-mass halo entered the horizon at
, that corresponding
to our cluster-mass halo at ten times this scale factor. Note that
both entered the horizon in the radiation-dominated epoch,
a ≪ a eq , so that indeed (
4.76 ) applies.

(E.4)


Solution to
7.4.
(1)
(2)
To obtain the next-order term, we write for a
flat universe
where we performed a Taylor expansion of the integrand in the
second step. A similar Taylor expansion of the Hubble function
yields
Multiplying these two results then yields
If we consider a significant deviation from the linear behavior to
occur when the second term in the parenthesis becomes of order 0.1,
then we request
. Taken the parameters which apply to our Universe,
, this becomes a ≲ 0. 6.
This estimate is in concordance with the behavior of the dashed
curve in Fig. 7.3 .
![$$\displaystyle\begin{array}{rcl} I(a)&:=& \int _{0}^{a} \frac{\mathrm{d}a^{{\prime}}} {\left (\varOmega _{\mathrm{m}}/a^{{\prime}} +\varOmega _{\varLambda }{a^{{\prime}}}^{2}\right )^{3/2}} {}\\ & =& \frac{1} {\varOmega _{\mathrm{m}}^{3/2}}\int _{0}^{a} \frac{\mathrm{d}a^{{\prime}}\;{a^{{\prime}}}^{3/2}} {\left [1 + (\varOmega _{\varLambda }/\varOmega _{\mathrm{m}}){a^{{\prime}}}^{3}\right ]^{3/2}} {}\\ & \approx & \frac{1} {\varOmega _{\mathrm{m}}^{3/2}}\int _{0}^{a}\mathrm{d}a^{{\prime}}\;{a^{{\prime}}}^{3/2}\left (1 -\frac{3} {2} \frac{\varOmega _{\varLambda }} {\varOmega _{\mathrm{m}}}{a^{{\prime}}}^{3}\right ) {}\\ & =& \frac{1} {\varOmega _{\mathrm{m}}^{3/2}} \frac{2} {5}a^{5/2}\left (1 -\frac{15} {22} \frac{\varOmega _{\varLambda }} {\varOmega _{\mathrm{m}}}a^{3}\right )\;, {}\\ \end{array}$$](A129044_2_En_BookBackmatter_OnlinePDF_Equ59.gif)




Solution to
8.1.
(1)
(3)
We first define the number density contrast
In analogy to ( 7.27 ), the angular correlation
function w is defined as
We next define the redshift probability distribution p ( z ) of the galaxies, given in terms of
p x ( x 3 ) by p ( z ) d z = p x ( x 3 ) d x 3 , where the function
x 3 (
z ) is given in (
4.53 ); however, we do not need to
use this relation explicitly. Then,
where we made use of the definition ( 7.27 ) of the correlation function
ξ g , and where
the argument of ξ
g depends on the separation of the two points
characterized by the directions
and
and
redshifts z
i . For this, we
can either use the comoving separation (as was done during most of
this and the previous chapter), or use the proper separation—they
just differ by a factor (1 + z ). The correlation ξ g is assumed to be zero
unless the two galaxies are close in space. In particular, for a
non-zero correlation the two galaxies need to have a similar
redshift. We thus define
,
, where
is the mean redshift and we
assume that ξ g
vanishes unless
. Then approximating p (
z 1 ) ≈
p ( z ) ≈ p ( z 2 ), and replacing the
integration over z
1 and z
2 by one over z
and Δ z , we obtain (
8.17 ), where the separation in the
argument of ξ g
is valid provided the angular separation θ ≪ 1, so that we can use tan
θ ≈ θ .









Solution to
8.2. We have
, where we used ( 4.52 ). Since X-ray emission is a
two-body process, we have L
X ∝
n e 2
V , where V ∝ R 3 is the volume of the
region considered, and R
= D A
θ . The temperature of the
gas can be determined independent of the distance, and just merely
enters in the constant of proportionality. Combining the two
expression for L
X , we find


or
. The
estimated gas mass is M
gas ∝ n
e V ∝
D A
5∕2 . The total mass within θ as determined from ( 6.37 ) can be written as
Varying D A adds
a constant to ln R and thus
does not change the derivative; hence, M ∝ D A . Together with the
previous result, we find f
gas ∝ D
A 3∕2 .


Solution to
8.3. We start with the Friedmann equation written as (
4.14 ),
and subtract from it the same equation specialized to the current
epoch ( a = 1) to obtain
so that the constant K is
eliminated. With
and
,
this can be rewritten as
which reproduces ( 4.81 ), with F given by ( 4.82 ), or
where we used ( 8.49 ) but included the curvature
term. Since
, the final term cannot increase as
a → 0. Hence, as in the
case of a cosmological constant, for early times the radiation term
dominates and the argument from Sect. 4.5.2 remains unchanged.




![$$\displaystyle{a^{2}\,H^{2}(a)\left [1 -\varOmega _{ 0}(a)\right ] = H_{0}^{2}\left [1 -\varOmega _{ 0}(a = 1)\right ]\;,}$$](A129044_2_En_BookBackmatter_OnlinePDF_Equbs.gif)
![$$\displaystyle{F(a) = \left ( \frac{\varOmega _{\mathrm{r}}} {a^{2}} + \frac{\varOmega _{\mathrm{m}}} {a} + [1 -\varOmega _{0}(a = 1)] + \frac{\varOmega _{\mathrm{DE}}} {a^{(1+3w)}}\right )^{-1}\;,}$$](A129044_2_En_BookBackmatter_OnlinePDF_Equbt.gif)

Index
Abell catalog. See Clusters of galaxies
Abell radius
AB magnitudes
4000 Å-break
absorption coefficient
absorption lines in quasar spectra
classification
Lyman-a forest ( see Lyman-a forest)
metal systems
accelerated expansion of the
Universe
accretion
Bondi–Hoyle–Lyttleton accretion
cold vs . hot accretion onto dark matter
halos
efficiency
radiatively inefficient accretion
spherical accretion
tidal disruption event (TDE)
accretion disk
advection-dominated accretion flow
corona
geometrically thin, optically thick
accretion disk
temperature profile
viscosity
Acceleration of particles
acoustic peaks
active galactic nuclei
broad-band energy distribution
active galaxies
absorption lines ( see absorption lines in quasar
spectra)
accretion
quasar mode
radio mode
anisotropic emission
big blue bump (BBB)
binary AGNs
black hole
black hole mass
scaling relation
black hole spin
BL Lac objects
blazars
broad absorption lines (BAL)
broad emission lines
in polarized light
broad line region (BLR)
classification
in clusters of galaxies
Compton thick AGN
Eddington ratio
energy generation
host galaxy
ionization cone
jets ( see jets)
LINERs
luminosity function
narrow line region (NLR)
obscuring ‘torus’
OVV (optically violently variable)
polarization
QSO (quasi-stellar object)
composite spectra
radio-loud & radio-quiet
dichotomy
quasars
radio emission
radio galaxies
broad-line radio galaxies (BLRG)
narrow-line radio galaxies (NLRG)
radio lobes
relativistic iron line
Seyfert galaxies
soft X-ray excess
spectra
TeV radiation
Type 1 AGNs
Type 2 QSO
unified models
variability
wide angle tail sources
X-ray emission
X-ray reflection
X-ray selection
X-shaped radio sources
active optics
adaptive mesh refinement (AMR)
adaptive optics
Advanced Camera for Surveys (ACS)
age-metallicity relation
Age of the Universe
air shower
AKARI
ALMA (Atacama Large
Millimeter/sub-millimeter Array)
alpha elements
Andromeda galaxy (M31)
Anglo-Australian Telescope (AAT)
angular correlations of galaxies.
See Correlation
function
Angular-diameter distance
angular momentum barrier
Angular resolution
anisotropy parameter ß of stellar orbits
anthropic principle
anti-particles
APEX (Atacama Pathfinder Experiment)
Arecibo telescope
ASCA (Advanced Satellite for Cosmology and
Astrophysics)
astronomical unit
asymmetric drift
asymptotic giant branch (AGB)
Atacama Cosmology Telescope (ACT)
attenuation of ? rays
Australian Square Kilometre Array Pathfinder
(ASKAP)
Baade’s Window
background radiation
infrared background (CIB)
of ionizing photons
limits from ?-ray blazars
microwave background ( see Cosmic microwave background)
X-ray background (CXB)
bar
baryogenesis
baryon asymmetry
baryonic acoustic oscillations (BAOs),
354–357, 395, 397–399, 430,449
as ‘standard rod’
in Lyman-a forest
measurements
baryons
Baryon-to-photon ratio
beaming
Beppo-SAX
biasing
of dark matter halos
as a function of galaxy
color
luminosity
Big Bang
blackbody radiation
energy density
Black holes
in AGNs
binary systems and merging
demography
evolution in mass
formation and evolution
in the Galactic center
in galaxies
at high redshift
kinematic evidence
mass growth rate
mass in AGNs
radius of influence
recoil
scaling with galaxy properties
Schwarzschild radius
BL Lac objects. See Active galaxies
blue-cloud galaxies
bolometric magnitude. See Magnitude
Bonner Durchmusterung
BOOMERANG
bosons
bottom-up structure formation
boxiness in elliptical galaxies
BPT (Baldwin–Phillips–Terlevich)
diagram
bremsstrahlung
brightest cluster galaxy (BCG)
brightness of the night sky
broad absorption line QSO (BAL QSOs).
See Active galaxies
brown dwarfs
bulge
of the MilkyWay
Bullet cluster
nature of dark matter
Butcher–Oemler effect
Canada-France-Hawaii Telescope (CFHT)
Canada-France Redshift Survey (CFRS)
CANDLES survey
cannibalism in galaxies
CCD (charge-coupled device)
Center for Astrophysics (CfA)-Survey
Cepheids
as distance indicators
period-luminosity relation
Cerro Chajnantor Atacama Telescope
(CCAT)
Chandra Deep Fields
Chandra satellite
Chandrasekhar mass
characteristic luminosity L*
chemical evolution
Cherenkov radiation
Cherenkov telescopes
CLASH (Cluster Lensing And Supernova survey
with Hubble) survey
clustering length
clusters of galaxies
Abell catalog
Abell radius
baryon content
beta model
brightest cluster galaxy (BCG)
Bullet cluster
Butcher–Oemler effect
catalogs
classification
color-magnitude diagram
Coma cluster, 13, 279, 293 ( see Coma cluster of galaxies)
cool-core clusters
cooling flows
cooling time
mass cooling rate
core radius
as cosmological probes
dark matter
distance class
evolution effects
extremely massive clusters
feedback
galaxy distribution
galaxy luminosity function
galaxy population
gas-mass fraction
HIFLUGCS catalog
intergalactic stars
intracluster medium
large-scale structure
luminosity function
mass calibration by weak lensing
mass determination
mass function
mass-luminosity relation
mass-temperature relation
mass-to-light ratio
mass-velocity dispersion relation
maxBCG catalog
mergers
metallicity of ICM
morphology
near-infrared luminosity
normalization of the power spectrum
number density
numerical simulations
projection effects
radio relics
RCS surveys
richness class
scaling relations
redshift dependence
self-similar behavior
selection effects
sound-crossing time
statistical mass calibration
temperature profile
Virgo cluster
weak lensing mass profile
X-ray radiation
X-ray spectrum
Y-parameter
Zwicky catalog
COBE (Cosmic Background Explorer)
cold dark matter (CDM)
substructure
collisionless gas
color-color diagram
color excess
color filter
color index
color-magnitude diagram
color temperature
Coma cluster of galaxies
distance
Sunyaev–Zeldovich observation
comoving coordinates
comoving observers
completeness and purity of samples
Compton Gamma Ray Observatory (CGRO)
Compton scattering
inverse
Compton- y parameter
concentration index of the NFW
profile
confusion limit
continuity equation
in comoving coordinates
convection
convergence point
cool-core clusters
cooling diagram
cooling fronts
cooling function
cooling of gas
the role of molecular hydrogen
and star formation
cooling time
Copernicus satellite
correlation function
angular correlation
anisotropy
definition
homogeneity and isotropy
of galaxies
from pair counts
projected
related to biasing
slope
correlation length
of Lyman-break galaxies
Silk damping
Cosmic Lens All-Sky Survey (CLASS)
cosmic luminosity density
cosmic microwave background
acoustic peaks
baryonic acoustic oscillations
cosmic variance
dipole
discovery
fluctuations
dependence on cosmological
parameters
discovery
foreground emission
gravitational lensing
integrated Sachs–Wolfe effect (ISW)
measuring the anisotropy
origin
polarization
primary anisotropies
redshift evolution
Sachs–Wolfe effect
secondary anisotropies
Silk damping
spectrum
Sunyaev–Zeldovich effect
temperature
Thomson scattering after reionization
Cosmic Origins Spectrograph (COS)
cosmic rays
acceleration
energy density
GZK cut-off
ultra-high energy cosmic rays
(UHECRs)
cosmic shear. See Gravitational lensing
cosmic variance
cosmic web
cosmological constant
smallness
the ‘why now’ problem
cosmological parameters
consistencies and discrepancies
degeneracies
determination
standard ?CDM model
cosmological principle
cosmology
classification of cosmological
models
components of the Universe
curvature scalar
dark ages
density fluctuations
epoch of matter–radiation equality
expansion equation
expansion rate
homogeneous world models
Newtonian cosmology
radiation density of the Universe
structure formation
tensor fluctuations
COSMOS survey
3CR radio catalog
curvature of the Universe
dark energy
equation-of-state
Dark Energy Survey (DES)
dark matter
in clusters of galaxies
cold and hot dark matter
filaments
in galaxies
seen in ‘bullet clusters’
in the Universe
warm dark matter
dark matter halos
angular momentum
biasing
contraction of gas
cooling of gas
Einasto profile
gas infall
mass function
Navarro–Frenk–White profile
number density
shapes
spin parameter
stellar mass-halo mass relation
substructure
universal mass profile
virial radius
virial temperature
weak gravitational lensing
deceleration parameter q 0
declination
DEEP2 redshift survey
deflection of light. See gravitational lensing
density contrast
density fluctuations in the Universe
origin
density parameter
as a function of redshift
deuterium
primordial
in QSO absorption lines
de Vaucouleurs law
diffraction limit
Digitized Sky Survey (DSS)
diskiness in elliptical galaxies
distance determination
of extragalactic objects
within the MilkyWay
distance ladder
distance modulus
distances in cosmology
distances of visual binary stars
D n
-s relation
Doppler broadening
Doppler effect
Doppler factor
Doppler favoritism
Doppler shift
Doppler width
downsizing
drop-out technique. See Lyman-break technique
dust
extinction and reddening
gray dust
infrared emission
warm dust
dust-to-gas ratio
dwarf galaxies. See galaxies
dynamical friction
dynamical heating
dynamical instability of
<I>N</I>-body systems
dynamical pressure
early-type galaxies
ecliptic
Eddington accretion rate
Eddington luminosity
Eddington ratio
effective radius
<I>R</I>e
effective temperature
Effelsberg radio telescope
Einasto profile
Einstein–de Sitter model. See Universe
Einstein observatory
Einstein radius ?E. See gravitational lensing
elementary particle physics
beyond the Standard Model
elliptical galaxies. See galaxies
cored profile
emission coefficient
energy density of a radiation field
energy efficiency of nuclear fusion
epoch of matter–radiation equality
equation of radiative transfer
equatorial coordinates
equivalent width
eROSITA space mission
escape fraction of ionizing photons
Euclid space mission
Euler equation
in comoving coordinates
European Extremely Large Telescope
(E-ELT)
expansion rate. See cosmology
Extended Medium Sensitivity Survey
(EMSS)
extinction
coefficient
and reddening
extremely red object (ERO)
Faber–Jackson relation
Fanaroff–Riley classification
Faraday rotation
feedback
by AGNs
by supernovae
Fermi bubbles
Fermi Gamma-Ray Space Telescope
fermions
filaments
Fingers of God
fireball model. See gamma-ray bursts
flatfield
flatness problem
fluid approximation
flux
free-fall time
free-free radiation
free streaming
Freeman law
Friedmann equations
Friedmann–Lemaître model
fundamental plane
tilt
FUSE (Far Ultraviolet Spectroscopic
Explorer)
Gaia
Galactic center
black hole
distance
Fermi bubbles
flares
X-ray echos
Galactic coordinates
cylindrical
galactic fountain model
Galactic latitude
Galactic longitude
Galactic plane
Galactic poles
galactic winds
in Lyman-break galaxies
galaxies
bimodal color distribution
blue cloud
brightness profile
BzK selection
cD galaxies
relation to intracluster light
characteristic luminosity
<I>L</I>*
chemical evolution
classification
color-color diagram
color-density relation
color-magnitude relation
color-profile shape relation
dark matter fraction
dark matter halo masses
distant red galaxies (DRGs)
dwarf galaxies
E+A galaxies
early-type galaxies
elliptical galaxies
blue compact dwarfs (BCD’s)
classification
composition
cores and extra light
counter-rotating disks
dark matter
dust lane
dwarf ellipticals (dE’s)
dwarf spheroidals (dSph’s)
dynamics
formation
gas and dust
indicators for complex evolution
interstellar medium
mass determination
mass fundamental plane
mass-to-light ratio
shape of the mass distribution
shells and ripples
star formation
stellar orbits
UV-excess
green valley
halos
at high redshift
high-redshift galaxies
color-magnitude distribution
demographics
interstellar medium
metallicity
mid-IR luminosity function
morphology
optical/NIR luminosity function
size and shape
size evolution
UV luminosity function
Hubble sequence
interacting galaxies
IRAS galaxies
irregular galaxies
late-type galaxies
LIRG (luminous infrared galaxy)
low surface brightness galaxies
(LSBs)
luminosity function
Lyman-a emitters (LAEs)
Lyman-break galaxies
correlation length
mass function
mass-metallicity relation
mass profile
mean number density
morphological classification
morphology-density relation
morphology of faint galaxies
narrow-band selection
polar ring galaxies
post-starburst galaxies
red sequence
S0 galaxies
gas and dust
satellite galaxies
scaling relations
spectra
spheroidal component
spiral galaxies
bars
bulge
bulges vs . pseudobulges
bulge-to-disk ratio
central surface brightness
color gradient
corona
dark matter
dust
dust obscuration and transparency
early-type spirals
gaseous halo
gas mass fraction
halo size
maximum disk model
metallicity
normal and barred
reddening
rotation curve
spiral structure
stellar halo
stellar populations
thick disk
warps
starburst galaxies
sub-millimeter galaxies
AGN fraction
correlation length
halo mass
identification in other wavebands
mergers
number counts
redshift distribution
substructure
suppression of low-mass galaxies
ULIRG (ultra-luminous infrared galaxy)
galaxy evolution
numerical simulations
overcooling problem
quasar epoch
semi-analytic models
Galaxy Evolution from Morphology and
Spectral Energy Distributions
(GEMS) survey
galaxy formation
formation of disk galaxies
overview
scale length of disks
galaxy groups
compact groups
diffuse optical light
Galaxy Zoo
GALEX (Galaxy Evolution Explorer)
gamma-ray bursts
afterglows
fireball model
hypernovae
short- and long-duration bursts
gauge bosons
G-dwarf problem
Gemini telescopes
General Relativity
Giant Magellan Telescope (GMT)
globular clusters
specific abundance
gluons
Gran Telescopio Canarias (GTC)
gravitational instability
gravitational lensing
AGN microlensing
clusters of galaxies as lenses
correlated distortions
of cosmic microwave background
cosmic shear
critical surface mass density
deflection angle
differential deflection
Einstein radius
Einstein ring
galaxies as lenses
galaxy-galaxy lensing
Hubble constant
lens equation
luminous arcs
magnification
mass determination
mass profile of dark matter halos
mass-sheet degeneracy
microlensing effect
microlensing magnification pattern
multiple images
point-mass lenses
search for clusters of galaxies
shear
shear correlation function
singular isothermal sphere (SIS)
model
substructure
time delay
weak lensing effect
gravitational redshift
gravitational waves
Great Attractor
Great Debate
Great Observatories Origins Deep Survey
(GOODS)
Great Wall
Green Bank Telescope
green-valley galaxies
groups of galaxies
growth factor <I>D</I>C
growth of density fluctuations
Gunn–Peterson effect
Gunn–Peterson test
near-zone transmission
HII-region
hadrons
harassment in galaxies
Harrison–Zeldovich fluctuation
spectrum
HEAO-1
heliocentric velocity
helium abundance
Herschel blank-field surveys
Herschel Lensing Survey (HLS)
Herschel Space Observatory
Hertzsprung–Russell diagram (HRD)
H.E.S.S. (High Energy Stereoscopic
System)
Hickson compact groups
hierarchical structure formation
Higgs mechanism
Higgs particle
highest-redshift objects
high-redshift galaxies
size evolution
high-velocity clouds (HVCs)
in external galaxies
Hipparcos
Hobby–Eberly Telescope
Holmberg effect
horizon
horizon length
at matter-radiation equality
horizon problem
hot dark matter (HDM)
Hubble classification of galaxies
Hubble constant <I>H</I>0
scaled Hubble constant h
Hubble Deep Field(s)
galaxy number counts
Hubble diagram
of supernovae
Hubble eXtreme Deep Field (XDF)
Hubble law
Hubble radius
Hubble sequence. See galaxies
Hubble Space Telescope (HST)
Hubble time
Hubble Ultradeep Field (HUDF)
hydrodynamics
hypernovae
hypervelocity stars
individual objects
? Centauri
3C 48
3C 75 = NGC1128
3C120
3C175
3C273
3C279
3C326
Abell 68
Abell 222 & 223
Abell 370
Abell 383
Abell 400
Abell 851
Abell 1689
Abell 1835
Abell 2218
Abell 2319
Abell 2597
Abell 3627
Andromeda galaxy (M31)
Antennae galaxies
Antlia dwarf galaxy
APM 08279+5255
Arp 148
Arp 220
Arp 256
B1938+666
B2045+265
BL Lacertae
Bullet cluster 1E 0657–56
Cartwheel galaxy
cB 58
Centaurus A = NGC 5128
Centaurus cluster
Centaurus group
CID-42
CIZA J2242.8+5301
Cl 0024+17
Cl 2244–02
Cl 0053–37
Cosmic Eye
Cygnus A
DLSCL J0916.2+2951
Dwingeloo 1
ESO 77-14
ESO 593-8
F10214+47
Fornax dwarf spheroidal
HCG40
HCG62
HCG87
Hercules A
Hercules cluster = Abell 2151
HFLS3
HXMM01
Hydra A
IRAS 13225–3809
IRAS 13349+2438
Leo I
M33
M51
M81
M81 group
M82
M83
M84
M86
M87
M94
MACS J0025.4–1222
MACS J0647+7015
MACS J0717.5+3745
MACS J1206.2_0847
MCG-6-30-15
MG 1654+13
MG2016+112
MS 0735.6+7421
MS 1054–03
MS 1512+36
NGC 17
NGC 253
NGC 454
NGC 474
NGC 1068
NGC 1232
NGC1265
NGC1275
NGC 1365
NGC 1705
NGC 2207 and IC 2163
NGC 2997
NGC 3115
NGC 3190
NGC 3198 (rotation curve)
NGC 4013
NGC 4151
NGC 4258
NGC 4261
NGC4402
NGC4522
NGC 4565
NGC 4631
NGC 4650A
NGC 5195
NGC 5548
NGC 5728
NGC 5866
NGC 5907
NGC 6050
NGC 6217
NGC 6240
NGC 6251
NGC 6670
NGC 6786
NGC6822 (Barnard’s Galaxy)
OJ 287
Perseus cluster
Pinwheel galaxy (M83)
PKS 1127_145
PKS 2155_304
PKS 2349
PSR J1915+1606
QSO 0957+561
QSO 1422+231
QSO 2237+0305
QSO PG1115+080
QSO ULAS J1120+0641
RXJ 1347–1145
Sculptor Group
SDSS J1030+0524
SDSS J1148+5251
Seyfert’s Sextet
SMM J09429+4658
SN 1987A
SPT-CL J2106–5844
Stephan’s Quintet (HCG92)
Tadpole galaxy = Arp 188
TN J1338–1942
UDFy-38135539
UGC 8335
UGC 9618
Whirlpool Galaxy (M51)
XMMUJ2235.2_2557
inflation
initial mass function (IMF)
integral field spectroscopy
Integral satellite
integrated Sachs–Wolfe effect (ISW)
interactions of galaxies
interferometry
intergalactic medium
intermediate-velocity clouds (IVCs)
interstellar medium
phases
intracluster light, 291, 330. See clusters of galaxies
intrinsic alignments of galaxies
inverse Compton scattering, See Compton scattering
ionization parameter
IRAS (InfraRed Astronomical Satellite)
IRAS galaxy surveys
irregular galaxies. See galaxies
ISO (Infrared Space Observatory)
isochrones
isophote
isothermal sphere
IUE (International Ultraviolet
Explorer)
James Webb Space Telescope (JWST)
Jansky (flux unit)
JCMT (James Clerk Maxwell Telescope)
Jeans equation
Jeans mass
jets
beaming
Doppler favoritism
generation and collimation
at high frequencies
K-correction
Keck telescope
Kilo Degree Survey (KiDS)
King models
Kirchhoff’s law
Kormendy relation
Kuiper Airborne Observatory
Large Binocular Telescope (LBT)
Large Hadron Collider (LHC)
Large Millimeter Telescope (LMT)
large-scale structure of the
Universe
baryon distribution
galaxy distribution
halo model
non-linear evolution
numerical simulations
Aquila Comparison Project
friends-of-friends algorithm
Hubble Volume Simulation
inclusion of feedback processes
Millennium Simulations
Virgo Simulation
power spectrum
tilt
Large Synoptic Survey Telescope (LSST)
Las Campanas Redshift Survey (LCRS)
laser guide star
Laser Interferometer Space Antenna
(LISA)
last-scattering surface
late-type galaxies
Leiden-Argentine-Bonn (LAB) survey
lenticular galaxies
leptons
light cone
light pollution
Limber equation
linearly extrapolated density fluctuation
field
linearly extrapolated power spectrum
LINERs, See active galaxies
line transitions: allowed, forbidden,
semi-forbidden
Local Group
galaxy content
mass estimate
local standard of rest (LSR)
look-back time
Lorentz factor
Low-Frequency Array (LOFAR)
luminosity
bolometric
in a filter band
luminosity classes
luminosity distance
luminosity function
evolution
of galaxies
of quasars
UV LF of high-redshift galaxies
luminous arcs. See gravitational lensing
luminous red galaxies (LRGs)
Lyman-a blobs
Lyman-a emitters (LAEs)
Lyman-a forest
baryonic acoustic oscillations
damped Lya systems
Lyman-limit systems
models
power spectrum
proximity effect
as a tool for cosmology
Lyman-break analogs
Lyman-break galaxies
seegalaxies
Lyman-break method
MACHOs
Madau diagram
Magellanic Clouds
distance
Magellanic Stream
MAGIC (Major Atmospheric Gamma-ray Imaging
Cherenkov
Telescopes)
magnification. See gravitational lensing
magnitude
absolute magnitude
apparent magnitude
bolometric magnitude
main sequence
Malmquist bias
MAMBO (Max-Planck Millimeter
Bolometer)
mass-energy equivalence
maser
mass segregation
mass spectrum of dark matter halos
mass-to-light ratio
in clusters of galaxies
of galaxies
maxBCG group catalog
MAXIMA
MeerKAT
merger tree
merging of galaxies
brightness profiles of merger
remnants
dry mergers
impact of AGN feedback
major merger
minor merger
wet mergers
MERLIN (Multi-Element Radio Linked
Interferometer Network)
mesons
metallicity
metallicity index
metals
microlensing. See gravitational lensing
MilkyWay
annihilation radiation
bar
bulge
center ( see Galactic center)
chemical composition
dark halo
disk
distribution of dust
gamma radiation
gas
gaseous halo
halo
hypervelocity stars
infalling gas
kinematics
magnetic field
multi-wavelength view
rotation curve
stellar streams
structure
thick disk
thin disk
Millennium Simulation
mixed dark matter (MDM)
Modified Newtonian Dynamics (MOND)
molecular clouds
moving cluster parallax
multi-object spectroscopy
narrow-band photometry
natural telescopes
Navarro–Frenk–White profile
Near Infrared Camera and Multi Object
Spectrograph (NICMOS)
neutrinos
freeze-out
masses
neutrino oscillations
radiation component of the Universe
Solar neutrino problem
Solar neutrinos
neutron stars
New General Catalog (NGC)
New Technology Telescope (NTT)
non-linear mass-scale
NVSS (NRAO VLA Sky Survey)
oblate and prolate
Olbers’ paradox
Oort constants
open clusters
optical depth
optically violently variables (OVV).
See active galaxies
outflows from galaxies
overcooling problem
pair production
and annihilation
Palomar Observatory Sky Survey (POSS)
PanSTARRS
parsec
particle-mesh (PM) method
particle-particle particle-mesh (P3M)
method
passive evolution of a stellar
population
Pauli exclusion principle
peak-background split
peculiar motion
peculiar velocity
period-luminosity relation
perturbation theory
photometric redshift
catastrophic outliers
Pico Veleta telescope
Pierre Auger Observatory
Planck function
Planck mass
Planck satellite
planetary nebulae
as distance indicators
Plateau de Bure interferometer
point-spread function
Poisson equation
in comoving coordinates
polarization
population III stars
population synthesis
power spectrum
of galaxies
normalization
shape parameter G
Press–Schechter model
pressure of radiation
primordial nucleosynthesis (BBN)
baryon density in the Universe
primordial spectral index n s
proper motion
proto-cluster
proximity effect
pseudobulges
pulsating stars
QSOs. See active galaxies
quarks
quasars. See active galaxies
radial velocity
radiation force
radiative transfer equation
RadioAstron
radio galaxies. See active galaxies
ram-pressure stripping
random field
ergodicity
Gaussian random field
realization
Rayleigh–Jeans approximation
reaction rate
recombination
two-photon decay
reddening vector
red cluster sequence (RCS)
red giants
red-sequence galaxies
redshift
cosmological redshift
desert
relation to scale factor
space
redshift space distortions
anisotropy of the correlation
function
redshift surveys of galaxies
reionization
helium reionization
observational probes
UV-slope of high-redshift galaxies
relaxation time-scale
reverberation mapping
right ascension
ROSAT (ROentgen SATellite)
ROSAT All-Sky Survey (RASS)
rotational flattening
rotation measure
RR Lyrae stars
runaway stars
Sachs–Wolfe effect
Sagittarius dwarf galaxy
Saha equation
Salpeter initial mass function. See initial mass function (IMF)
satellite galaxies. See galaxies
scale factor. See Universe
scale-height of the Galactic disk
Schechter luminosity function
Schmidt–Kennicutt law of star
formation
SCUBA (Sub-millimeter Common-User Bolometer
Array)
SCUBA-2
SDSS Quasar Lens Search (SQLS)
secondary distance indicators
seeing
self-shielding
semi-analytic model of galaxy
evolution
sensitivity of telescopes
Sérsic brightness profile
Sérsic index
service mode observing
Seyfert galaxies. See Active galaxies
SgrA_. See Galactic center
shape parameter G
shells and ripples
shock fronts
shock heating
Silk damping
singular isothermal sphere (SIS)
Sloan Digital Sky Survey (SDSS)
Sloan Great Wall
Sloan Lens Advanced Camera for Surveys
(SLACS)
smooth particle hydrodynamics (SPH)
SOFIA (Stratospheric Observatory for
Infrared Astronomy)
softening length
sound horizon
sound velocity in photon-baryon
fluid
sound waves
source counts in an Euclidean
universe
source function
South Pole Telescope (SPT)
Space Telescope Imaging Spectrograph
(STIS)
specific energy density of a radiation
field
specific intensity
spectral classes
spectral resolution
spectroscopic distance
spherical accretion
spherical collapse model
spin parameter
spiral arms
as density waves
spiral galaxies. See galaxies
Spitzer Space Telescope
Square Kilometer Array (SKA)
standard candles
starburst-AGN connection
starburst galaxies. See galaxies
star formation
and color of galaxies
comparison of indicators
cosmic history
different modes
efficiency
feedback processes
initial mass function (IMF)
Madau diagram
quiescent star formation
rate
Schmidt–Kennicutt law
star-formation burst
star-formation rate (SFR)
star-formation rate density
Stefan–Boltzmann law
stellar evolution
stellar mass estimate
stellar mass function
stellar populations
stellar streams
strangulation in galaxies
strong interaction
Subaru telescope
Sunyaev–Zeldovich effect
Compton- y parameter
distance determination
Hubble constant
integrated Compton- y parameter
kinetic SZ effect
superclusters
superluminal motion
supernovae
classification
core-collapse supernovae
as distance indicators
evolutionary effects
at high redshift
metal enrichment of the ISM
SN 1987A
Type Ia
single- vs . double-degenerate progenitor
supersymmetry
surface brightness fluctuations
surface gravity
surveys
Suzaku
SWIFT
synchrotron radiation
cooling time of electrons
polarization
spectral shape
synchrotron self-absorption
synchrotron self-Compton radiation
tangential velocity
tangent point method
TeV-astronomy
thermal radiation
Thirty Meter Telescope (TMT)
Thomson cross section
Thomson scattering
three-body dynamical system
throughput
tidal disruption
tidal disruption event (TDE)
tidal streams
tidal stripping
tidal tails
tidal torque
time dilation
tip of red giant branch
transfer function
qualitative behavior
triaxial ellipsoid
trigonometric parallax
of Cepheids
Tully–Fisher relation
baryonic
Two-Degree-Field Survey
Two Micron All Sky Survey (2MASS)
ultra-luminous compact X-ray sources
(ULXs)
ultra-luminous infrared galaxies (ULIRGs).
See galaxies
Universe
age
baryon asymmetry
baryon-to-photon ratio
critical density
as a function of redshift
curvature
density
density fluctuations
density parameter
Einstein–de Sitter model
expansion
homogeneity scale
scale factor
standard model
thermal history
vacuum energy. See dark energy
variation of physical constants
velocity dispersion
in clusters of galaxies
in galaxies
Very Large Array (VLA)
Very Large Telescope (VLT)
Very Long Baseline Array (VLBA)
Very Long Baseline Interferometry
(VLBI)
VIMOS VLT Deep Survey (VVDS)
violent relaxation
Virgo cluster of galaxies
intracluster light
virial mass
virial radius
virial temperature
virial theorem
virial velocity
virtual observatory
VISTA
VLT Survey Telescope (VST)
voids
Voigt profile
W- and Z-boson
warm dark matter
warm-hot intergalactic medium
wave number
weak interaction
weak lensing effect. See gravitational lensing
wedge diagram
white dwarfs
Wide Field and Planetary Camera
(WFPC2)
Wide Field Camera 3 (WFC3)
width of a spectral line
Wien approximation
Wien’s law
WIMPs
direct detection experiments
indirect astrophysical signatures
WISE (Wide-field Infrared Survey
Explorer)
WMAP (Wilkinson Microwave Anisotropy
Probe)
Wolter telescope
X-factor
XMM-Newton
X-ray absorption
X-ray background
X-ray binaries
X-ray source counts
yield
Zeeman effect
zero-age main sequence (ZAMS)
zodiacal light
Zone of Avoidance
Footnotes
1
Of course, this convention is confusing,
particularly to someone just becoming familiar with astronomy, and
it frequently causes confusion and errors, as well as problems in
the communication with non-astronomers—but we have to get along
with that.
2
The detection of neutrinos from the Sun in
terrestrial detectors was the final proof for the energy production
mechanism being nuclear fusion.
3
If the mass of a brown dwarf exceeds ∼ 0. 013
M ⊙ , the
central density and temperature are high enough to enable the
fusion of deuterium (heavy hydrogen) into helium. However, the
abundance of deuterium is smaller by several orders of magnitude
than that of normal hydrogen, rendering the fuel reservoir of a
brown dwarf very small.