© Springer-Verlag Berlin Heidelberg 2015
Peter SchneiderExtragalactic Astronomy and Cosmology10.1007/978-3-642-54083-7_9

9. The Universe at high redshift

Peter Schneider
(1)
Argelander-Institut für Astronomie, Universität Bonn, Bonn, Germany
 
In the previous chapter we explained by what means the cosmological parameters may be determined, and what progress has been achieved in recent years. This might have given the impression that, with the determination of the values for Ω m, Ω Λ etc., cosmology is nearing its conclusion. As a matter of fact, for several decades cosmologists have considered the determination of the density parameter and the expansion rate of the Universe as their prime task, and now this goal has largely been achieved. However, from this point on, the future evolution of the field of cosmology will probably proceed in two directions. First, we will try to uncover the nature of dark energy and to gain new insights into fundamental physics along the way. Second, astrophysical cosmology is much more than the mere determination of a few parameters. We want to understand how the Universe evolved from a very primitive initial state, as seen in the almost isotropic CMB radiation, into what we are observing around us today—galaxies of different morphologies, luminosities and spectral properties, the large-scale structure of their distribution, groups and clusters of galaxies, active galaxies, and the intergalactic medium. We seek to study the formation of stars and of metals, the cosmic history of star formation, and also the processes that reionized and enriched the intergalactic medium with heavy elements.
The boundary conditions for studying these processes are now very well defined. Until about the year 2000, the cosmological parameters in models of galaxy evolution, for instance, could be chosen from within a large range, because they had not been determined sufficiently well at that time. That allowed these models more freedom to adjust the model outcomes such that they best fit with observations. Today however, a successful model needs to make predictions compatible with observations, but using the parameters of the standard model. In terms of the cosmological parameters, there is little freedom left in designing such models. In other words, the stage on which the formation and evolution of objects and structure takes place is prepared, and now the cosmic play can begin.
/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig1_HTML.jpg
Fig. 9.1
Spectrum of a QSO at the high redshift of z = 6. 419. Like many other QSOs at very high redshift, this source was discovered with the Sloan Digital Sky Survey. The spectrum was obtained with the Keck telescope. The redshifted Lyα line is clearly visible, its blue side ‘eaten’ away by intergalactic absorption. Almost all radiation bluewards of the central wavelength of the Lyα line is absorbed; however, a low level of this radiation is getting through, as is most clearly seen from the Lyβ line. For λ ≤ 7200 Å the spectral flux is consistent with zero; intergalactic absorption is too strong here. Source: X. Fan et al. 2003, A Survey of z > 5.7 Quasars in the Sloan Digital Sky Survey. II. Discovery of Three Additional Quasars at z > 6, AJ 125, 1649, p. 16554, Fig. 6. ©AAS. Reproduced with permission
Progress in recent years, with developments in instrumentation having played a vital role, has allowed us to examine the Universe at very high redshift. An obvious indication of this progress is the increasingly high maximum redshift of sources that can be observed; as an example, Fig. 9.1 presents the spectrum of a QSO at redshift z = 6. 419 whose precise redshift was measured from molecular CO lines. Today, we know quite a few galaxies at redshift z > 6, i.e., we observe these objects at a time when the Universe had less than 10 % of its current age and when the density of the neutral hydrogen in the intergalactic medium was apparently considerably higher than at later epochs, as concluded from the very strong absorption blueward of the Lyα emission line (see Fig. 9.1). As we shall see, the detection of galaxies at even higher redshifts has been claimed. Besides larger telescopes, which enabled these deep images of the Universe, gaining access to new wavelength domains is of particular importance for our studies of the distant Universe. This can be seen, for example, from the fact that the optical radiation of a source at redshift z ∼ 1 is shifted into the NIR. Because of this, near-infrared astronomy is about as important for galaxies at z ≳ 1 as optical astronomy is for the local Universe. Furthermore, the development of sub-millimeter astronomy has provided us with a view of sources that are nearly completely hidden to the optical eye because of strong dust absorption.
In this chapter, we will attempt to provide an impression of astronomy of the distant Universe, and shed light on some interesting aspects that are of particular importance for our understanding of the evolution of the Universe, whereas in Chap. 10, we will try to provide an impression of our theoretical understanding of the evolution of galaxies throughout the Universe. Both, observational as well as theoretical and numerical studies, are currently very rapidly developing fields of research, so we will simply address some of the main topics in this field today. We begin in Sect. 9.1 with a discussion of methods to specifically search for high-redshift galaxies, and we will then focus on a method by which galaxy redshifts can be determined solely from photometric information in several bands (thus, from the color of these objects). This method can be applied to deep multi-band sky images, and we will present some of the results from deep HST surveys, described in Sect. 9.2. We will also emphasize the importance of gravitational lenses as ‘natural telescopes’, which provide us with a deeper view into the Universe due to their magnification effect.
Gaining access to new wavelength domains paves the way for the discovery of new kinds of sources; in Sect. 9.3 we will present high-redshift galaxy populations, some of which have been identified by sub-millimeter and NIR observations. Some key properties of the high-redshift galaxy population will be described in Sect. 9.4, including their luminosity function; as will be shown there, the properties of galaxies in the early phases of our Universe are quite different from the present galaxies. In Sect. 9.5 we will show that, besides the CMB, background radiation also exists at other wavelengths, but whose nature is considerably different from that of the CMB; recent progress has allowed us to identify the nature of these cosmic backgrounds. Then, in Sect. 9.6, we will focus on the history of cosmic star formation, and show that at redshift z ≳ 1 the Universe was much more active than it is today—in fact, most of the stars that are observed in the Universe today were already formed in the first half of cosmic history. This empirical discovery is one of the aspects that one attempts to explain in the framework of models of galaxy formation and evolution. Finally, in Sect. 9.7 we will discuss the sources of gamma-ray bursts. These are explosive events which, for a very short time, appear brighter than all other sources of gamma rays on the sky put together. For about 25 years the nature of these sources was totally unknown; even their distance estimates were spread over at least seven orders of magnitude. Only since 1997 has it been known that these sources are of extragalactic origin.

9.1 Galaxies at high redshift

In this section we will first consider how distant galaxies can be found, and how to identify them as such. The properties of these high-redshift galaxies can then be compared with those of galaxies in the local Universe, which were described in Chap. 3. The question then arises as to whether galaxies at high z, and thus in the early Universe, look like local galaxies, or whether their properties are completely different. One might, for instance, expect that the mass and luminosity of galaxies are evolving with redshift since, as we have seen in Sect. 7.​5.​2, the mass function of dark matter halos changes during cosmic evolution. Examining the galaxy population as a function of redshift, one can trace the history of global cosmic star formation and analyze when most of the stars visible today have formed, and how the density of galaxies changes as a function of redshift. We will investigate some of these questions in this and the following sections.
How to find high-redshift galaxies? Until about 1995 only a few galaxies with z > 1 had been known; most of them were radio galaxies discovered by optical identification of radio sources. The most distant normal galaxy with z > 2 then was the source of the giant luminous arc in the galaxy cluster Cl 2244−02 (see Fig. 6.​49). Very distant galaxies are expected to be faint, and so the question arises of how galaxies at high z can be detected at all.
The most obvious answer to this question may perhaps be by spectroscopy of a sample of faint galaxies. This method is not feasible though, since galaxies with R ≲ 22 have redshifts z ≲ 0. 5, and spectra of galaxies with R > 22 are observable only with 4-m telescopes and with a very large investment of observing time.1 Also, the problem of finding a needle in a haystack arises: most galaxies with R ≲ 24. 5 have redshifts z ≲ 2 (a fact that was not established before 1995), so how can we detect the small fraction of galaxies with larger redshifts?
Narrow-band photometry. A more systematic method that has been applied is narrow-band photometry. Since hydrogen is the most abundant element in the Universe, one expects that some fraction of galaxies feature a Lyα emission line (as do all QSOs). By comparing two sky images, one taken with a narrow-band filter centered on a wavelength λ, the other with a broader filter also centered roughly on λ, this line emission can be searched for specifically. If a galaxy at  $$z\,\approx \,\lambda /(1216\,\mbox{ \AA }) - 1$$ has a strong Lyα emission line, it should be particularly bright in the narrow-band image in comparison to the broad-band image, relative to other sources. This search for Lyα emission line galaxies had been almost without success until the mid-1990s. Among other reasons, one did not know what to expect, e.g., how faint galaxies at z ≳ 3 are and how strong their Lyα line would be. Another reason, which was found only later, was the leakage of the narrow-band filters for radiation at shorter and longer wavelength—the transmission of these filters was not close enough to zero for wavelengths outside the considered narrow range. We will see later that more recent narrow-band photometric surveys have indeed uncovered a population of high-redshift galaxies.

9.1.1 Lyman-break galaxies (LBGs)

/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig2_HTML.jpg
Fig. 9.2
Principle of the Lyman-break method. The histogram shows the synthetic spectrum of a galaxy at z = 3. 15, generated by models of population synthesis; the spectrum belongs to a QSO at slightly higher redshift. Clearly, the decline of the spectrum at λ ≤ 912(1 + z)Å is noticeable. Furthermore, we see that the flux for λ ≤ 1216(1 + z)Å is reduced relative to the radiation on the red side of the Lyman-α emission line due to the integrated absorption of the intergalactic Lyman-α forest. For higher redshift sources, this latter effect becomes stronger, so that for them the break occurs already at a rest wavelength of λ = 1216 Å. The three dotted curves are the transmission curves of three broad-band filters, chosen such that one of them (Un) blocks all photons with wavelengths above the Lyman-break. The color of this galaxy would then be blue in  $$\mathrm{G} -\mathcal{R}$$ , and very red in Un − G. Source: C.C. Steidel et al. 1995, Lyman Imaging of High-Redshift Galaxies. III. New Observations of Four QSO Fields, AJ 110, 2519, p. 2520, Fig. 1. ©AAS. Reproduced with permission
A129044_2_En_9_Fig3_HTML.gif
Fig. 9.3
Top panel: A U-band drop-out galaxy. It is clearly detected in the two redder filters, but vanishes almost completely in the U-filter. Bottom panel: In a single CCD frame, a large number of candidate Lyman-break galaxies ( ∼ 150) are found. They are marked with circles here; their density is about 1 per square arcminute. Credit: C.C. Steidel
The method. The breakthrough was obtained with a method that became known as the Lyman-break method. Since hydrogen is so abundant and its ionization cross section so large, one can expect that photons with λ < 912 Å are very heavily absorbed by neutral hydrogen in its ground state. Therefore, photons with λ < 912 Å have a low probability of escaping from a galaxy without being absorbed.
Intergalactic absorption also contributes. In Sect. 5.​7 we saw that each QSO spectrum features a Lyα forest and often also Lyman-limit absorption. The intergalactic gas absorbs a large fraction of photons emitted by a high-redshift source at λ < 1216 Å, and virtually all photons with a rest-frame wavelength λ ≲ 912 Å. As also discussed in Sect. 8.​5.​2, the strength of this absorption increases with increasing redshift. Combining these facts, we conclude that spectra of high-redshift galaxies should display a distinct feature—a ‘break’—at λ = 912 Å for redshifts z ≲ 4, and for higher redshifts, the break shifts more towards λ = 1216 Å. Furthermore, radiation with λ ≲ 912 Å should be strongly suppressed by intergalactic absorption, as well as by absorption in the interstellar medium of the galaxies themselves, so that only a very small fraction of these ionizing photons will reach us.
From this, a strategy for the detection of galaxies at z ≳ 3 emerges. We consider three broad-band filters with central wavelengths λ 1 < λ 2 < λ 3, where their spectral ranges are chosen to not (or only marginally) overlap. If λ 1 ≲ (1 + z)912 Å ≲ λ 2, a galaxy containing young stars should appear relatively blue as measured with the filters λ 2 and λ 3, and be virtually invisible in the λ 1-filter: because of the absorption, it will drop out of the λ 1-filter (see Fig. 9.2). For this reason, galaxies that have been detected in this way are called Lyman-break galaxies (LBGs) or drop-outs. An example of this is displayed in Fig. 9.3.
Large samples of LBGs. The method was first applied systematically in 1996, using the filters specified in Fig. 9.2. As can be read from Fig. 9.4, the expected location of a galaxy at z ∼ 3 in a color-color diagram with this set of filters is nearly independent of the type and star formation history of the galaxy. Hence, sources in the relevant region of the color-color diagram are very good candidates for being galaxies at z ∼ 3. The redshift needs to be verified spectroscopically, but the crucial point is that the color selection of candidates yields a very high success rate per observed spectrum, and thus spectroscopic observing time at the telescope is spent very efficiently in confirming the redshift of distant galaxies. With the commissioning of the Keck telescope (and later also of other telescopes of the 10-m class), spectroscopy of galaxies with B ≲ 25 became possible (see Fig. 9.5). Employing this method, thousands of galaxies with 2. 5 ≲ z ≲ 3. 5 have been detected and spectroscopically verified to date.
/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig4_HTML.jpg
Fig. 9.4
Left panel: Evolutionary tracks of galaxies in the ( $$\mathrm{G} -\mathcal{R}$$ ) – (Un − G) color-color diagram, for different types of galaxies, as obtained from population synthesis models. All evolutionary tracks start at z = 0, and the symbols along the curves mark intervals of Δ z = 0. 1. The colors of the various galaxy types are very different at lower redshift, but for z ≥ 2. 7, the evolutionary tracks for the different types nearly coincide—a consequence of the Lyα absorption in the intergalactic medium. Hence, a color selection of galaxies in the region between the dotted and dashed curves should select galaxies with z ≥ 3. Indeed, this selection of candidates has proven to be very successful; thousands of galaxies with z ∼ 3 have been spectroscopically verified. Right panel: The same color-color diagram, with objects selected from one survey field. The green and yellow shaded regions show the selection criteria for z ∼ 3 Lyman-break galaxies, the cyan and magenta regions indicate the selection windows for galaxies with z ∼ 2. 2 and z ∼ 1. 7, respectively. The symbols are coded according to the brightness of the sources, and triangles denote sources for which only lower limits in the Un − G color were obtained. Source: Left: C.C. Steidel et al. 1995, Lyman Imaging of High-Redshift Galaxies. III. New Observations of Four QSO Fields, AJ 110, 2519, p. 2522, Fig. 2. ©AAS. Reproduced with permission. Right: C.C. Steidel et al. 2004, A Survey of Star-forming Galaxies in the 1. 4 ≲ z ≲ 2. 5 Redshift Desert: Overview, ApJ 604, 534, p. 537, Fig. 1. ©AAS. Reproduced with permission
From the spectra shown in Fig. 9.5, it also becomes apparent that not all galaxies that fulfill the selection criteria also show a Lyα emission line, which provides one of the explanations for the lack of success in earlier searches for high-redshift galaxies using narrow-band filters. The spectra of the high-redshift galaxies which were found by this method are very similar to those of starburst galaxies at low redshift. It should come as no surprise that the galaxies selected by the drop-out technique feature active star formation, since it was required that the spectrum on the red side of the break—i.e., at (rest-frame) wavelengths above 1216 Å—shows a blue spectrum. Such a blue spectrum in the rest-frame UV is produced only by a stellar population which features active star formation. Furthermore, the luminosity of galaxies in the rest-frame UV and blue range strongly depends on the star-formation rate, so that preferentially galaxies with the highest (unobscured—see below) star-formation rate are selected.
This is a prominent example of the effect that the physical properties of objects selected depend on the selection criteria. One must always bear in mind that, when comparing galaxy populations detected by different methods, the properties can differ substantially. One of the challenges of studies of (high-redshift) galaxies is to get a coherent picture of the galaxy population from samples with a vast variety of selection methods.
The correlation function and halo masses of LBGs. For a large variety of objects, and over a broad range of separations, the spatial correlation function of objects can be described by the power law (7.​28), with a slope of typically γ ∼ 1. 7. However, the amplitude of this correlation function varies between different classes of objects. For example, we saw in Sect. 8.​2.​4 that the amplitude of the power spectrum of galaxy clusters is larger by about a factor 7 than that of galaxies (see Fig. 8.​23); the same ratio holds of course for the corresponding correlation functions. As we argued there, the strength of the correlation depends on the mass of objects; in the simple picture of biasing shown in Fig. 7.​22, the correlation of objects is larger the rarer they are. High-mass peaks exceeding the density threshold needed for gravitational collapse have a lower mean number density than low-mass peaks, so they are therefore expected to be more biased (see Sect. 8.​1.​3) and thus more strongly correlated.
If we now assume that each galaxy lives in a dark matter halo, we can estimate the dark halo mass from the observed correlation function of these galaxies. As we discussed in Sect. 7.​6.​3, dark matter halos have clustering properties which differ from the clustering of the underlying matter density field, and we described that in terms of the halo biasing b h(M, z), which is a function of halo mass and redshift. The dark matter correlation function can be determined quite accurately from numerical simulations. The ratio of the observed correlation function to the dark matter correlation function then yields the square of the halo bias parameter (7.​68), and comparing that to the numerically-determined function b h(M, z), the corresponding halo mass can be obtained.
/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig5_HTML.jpg
Fig. 9.5
Spectra of two galaxies at z ∼ 3, detected by means of the U-drop-out technique. Below each spectrum, the spectrum of a nearby starburst galaxy (NGC 4214)—shifted to the corresponding redshift—is plotted; it becomes apparent that the spectra of galaxies at z ∼ 3 are very similar to those of present-day star-forming galaxies. One of the two U-drop-out galaxies features a strong Lyα emission line, the other shows absorption at the respective wavelength. Source: C.C. Steidel et al. 1996, Spectroscopic Confirmation of a Population of Normal Star-forming Galaxies at Redshifts z > 3, ApJ 462, L17, PLATE L3, Fig. 1. ©AAS. Reproduced with permission
Considering the spatial distribution of LBGs, we find a large correlation amplitude. The (comoving) correlation length of LBGs at redshifts 1. 5 ≲ z ≲ 3. 5 is r 0 ∼ 4. 2h −1 Mpc, i.e., not very different from the correlation length of L -galaxies in the present Universe. Since the bias factor of present-day galaxies is about unity, implying that they are clustered in a similar way as the dark matter distribution, this result then implies that the bias of LBGs at high redshift must be considerably larger than unity. This conclusion is based on the fact that the dark matter correlation at high redshifts (on large scales, i.e., in the linear regime) was smaller than today by the factor D + 2(z), where D + is the growth factor of linear perturbations introduced in Sect. 7.​2.​2. Thus we conclude that LBGs are rare objects and thus correspond to high-mass dark matter halos. Comparing the observed correlation length r 0 with numerical simulations, the characteristic halo mass of LBGs can be determined, yielding ∼ 3 × 1011 M at redshifts z ∼ 3, and ∼ 1012 M at z ∼ 2. Furthermore, the correlation length is observed to increase with the luminosity of the LBG, indicating that more luminous galaxies are hosted by more massive halos, which are more strongly biased than less massive ones. If these results are combined with the observed correlation functions of galaxies in the local Universe and at z ∼ 1, and with the help of numerical simulations, then this indicates that a typical high-redshift LBG will evolve into a massive elliptical galaxy by today.
Proto-clusters. Furthermore, the clustering of LBGs shows that the large-scale galaxy distribution was already in place at high redshifts. In some fields the observed overdensity in angular position and galaxy redshift is so large that one presumably observes galaxies which will later assemble into a galaxy cluster—hence, we observe some kind of proto-cluster. We have already shown such a proto-cluster in Fig. 6.​71. Galaxies in such a proto-cluster environment seem to have about twice the stellar mass of those LBGs outside such structures, and the age of their stellar population appears older by a factor of two. This result indicates that the stellar evolution of galaxies in dense environments proceeds faster than in low-density regions, in accordance with expectations from structure formation. It also reveals a dependence of galaxy properties on the environment, which we have seen before manifested in the morphology-density relation (see Sect. 6.​7.​2). Proto-clusters of galaxies have also been detected at higher redshifts up to z ∼ 6, using narrow-band imaging searches for Lyman-alpha emission galaxies (see below).
Satellite galaxies at high redshifts. Whereas the clustering of LBGs is well described by the power law (7.​28) over a large range of scales, the correlation function exhibits a significant deviation from this power law at very small scales: the angular correlation function exceeds the extrapolation of the power law from larger angles at Δ θ ≲ 7″, corresponding to comoving length-scales of ∼ 200 kpc. It thus seems that this scale marks a transition in the distribution of galaxies. To get an idea of the physical nature of this transition, we note that this length-scale is about the virial radius of a dark matter halo with M ∼ 3 × 1011 M , i.e., the mass of halos which host the LBGs. On scales below this virial radius, the correlation function thus no longer describes the correlation between two distinct dark matter halos. An interpretation of this fact is provided in terms of merging: when two galaxies and their dark matter halos merge, the resulting dark matter halo hosts both galaxies, with the more massive one close to the center and the other one as ‘satellite galaxy’. The correlation function on scales below the virial radius thus indicates the clustering of galaxies within the same halo, whereas on larger scales, where it follows the power-law behavior, it indicates the correlation between different halos. Note that this effect is also well described in the halo model which we discussed in Sect. 7.​7.​3. On large scales, the correlation function is dominated by the two-halo term, whereas on smaller scales, the one-halo term takes over. The transition between these two regimes, which at low redshifts occurs on scales of several hundred kiloparsecs (see Fig. 7.​27), is at smaller scales for high-redshift galaxies, since the high-mass population of galaxy clusters has not formed yet at these early epochs.
Winds of star-forming galaxies. The inferred high star-formation rates of LBGs implies an accordingly high rate of supernova explosions. These release part of their energy in the form of kinetic energy to the interstellar medium in these galaxies. This process will have two consequences. First, the ISM in these galaxies will be heated locally, which slows down (or prevents) further star formation in these regions. This thus provides a feedback effect for star formation which prevents all the gas in a galaxy from turning into stars on a very short time-scale, and is essential for understanding the formation and evolution of galaxies, as we shall see in Sect. 10.​4.​4. Second, if the amount of energy transferred from the SNe to the ISM is large enough, a galactic wind may be launched which drives part of the ISM out of the galaxy into its halo. Evidence for such galactic winds has been found in nearby galaxies, for example from neutral hydrogen observations of edge-on spirals which show an extended gas distribution outside the disk. Furthermore, the X-ray corona of spirals (see Fig. 3.​26) is most likely linked to a galactic wind in these systems.
Indeed, there is now clear evidence for the presence of massive winds from LBGs. The spectra of LBGs often show strong absorption lines, e.g., of Civ, which are blueshifted relative to the velocity of the emission lines in the galaxy. An example of this effect can be seen in the spectra of Fig. 9.5, where in the upper panel the emission line of Civ is accompanied by an absorption to the short-wavelength side of the emission line. Such absorption can be produced by a wind moving out from the star-forming regions of the galaxy, so that its redshift is smaller than that of the emission regions. Characteristic velocities are ∼ 200 km∕s. In one case where the spectral investigation has been performed in most detail (the LBG cB58; see Fig. 9.17), the outflow velocity is ∼ 255 km∕s, and the outflowing mass rate exceeds the star-formation rate. Whereas these observations clearly show the presence of outflowing gas, it remains undetermined whether this is a fairly local phenomenon, restricted to the star-formation sites, or whether it affects the ISM of the whole galaxy.
Connection to QSO absorption lines. A slightly more indirect argument for the presence of strong winds from LBGs comes from correlating the absorption lines in background QSO spectra with the position of LBGs. These studies have shown that whenever the sight-line of a QSO passes within ∼ 40 kpc of an LBG, very strong Civ absorption lines (with column density exceeding 1014 cm−2) are produced, and that the corresponding absorbing material spans a velocity range of Δ v ≳ 250 km∕s; for about half of the cases with impact parameters within 80 kpc, strong Civ absorption is produced. This frequency of occurrence implies that about 1/3 of all Civ metal absorption lines with N ≳ 1014 cm−2 in QSO spectra are due to gas within ∼ 80 kpc from those LBGs which are bright enough to be included in current surveys. It is plausible that many of the remaining 2/3 are due to fainter LBGs.
The association of Civ absorption line systems with LBGs by itself does not prove the existence of winds in such galaxies; in fact, the absorbing material may be gas orbiting in the halo in which the corresponding LBG is embedded. In this case, no outflow phenomenon would be implied. However, in that case one might wonder where the large amount of metals implied by the QSO absorption lines is coming from. They could have been produced by an earlier epoch of star formation, but in that case the enriched material must have been expelled from its production site in order to be located in the outer part of z ∼ 3 halos. It appears more likely that the production of metals in QSO absorption systems is directly related to the ongoing star formation in the LBGs. We shall see in Sect. 9.3.5 that clear evidence for superwinds has been discovered in one massive star-forming galaxy at z ∼ 3.
A129044_2_En_9_Fig6_HTML.gif
Fig. 9.6
A galaxy at z = 5. 74, which is visible in the narrow-band filter (upper left panel) and in the I- and z-band (located between the two horizontal dashes), but which does not show any flux in the three filters at shorter wavelength (lower panels). Source: Hu et al. 1999, An Extremely Luminous Galaxy at z = 5. 74, ApJ 522, L9, p. L10, Fig. 1. ©AAS. Reproduced with permission
Finally, we mention another piece of evidence for the presence of superwinds in star-forming galaxies. There are indications that the density of absorption lines in the Lyα forest is reduced when the sight-line to the QSO passes near a foreground LBG. This may well be explained by a wind driven out from the LBG, pushing neutral gas away and thus leaving a gap in the Lyα forest. The characteristic size of the corresponding ‘bubbles’ is estimated to be ∼ 0. 5 Mpc for luminous LBGs.
Lyman-break galaxies at low redshifts. One might ask whether galaxies similar to the LBGs at z ∼ 3 exist in the current Universe. Until recently this question was difficult to investigate, since it requires imaging of lower redshift galaxies at ultraviolet wavelengths. With the launch of GALEX an appropriate observatory became available with which to observe galaxies with restframe UV luminosities similar to those of LBGs. UV-selected galaxies show a strong inverse correlation between the stellar mass and the surface brightness in the UV. Lower-mass galaxies are more compact than those of higher stellar mass. On the basis of this correlation we can consider the population of large and compact UV-selected galaxies separately. The larger ones show a star-formation rate of a few M ∕yr; at this rate, their stellar mass content can be built up on a time-scale comparable to the Hubble time, i.e., the age of the Universe. These galaxies are typically late-type spiral galaxies, and they show a metallicity similar to our Galaxy.
In contrast, the compact galaxies have a lower stellar mass and about the same star-formation rate, which allows them to generate their stellar population much faster, in about 1 Gyr. Compared to normal low-redshift galaxies, their metallicity is smaller by about a factor of 2 for a given stellar mass. In addition, they show similar extinction and outflow properties as the LBG at z ∼ 3. Hence, the properties of the compact UV-selected galaxies, which are sometimes called Lyman-break analogs , are quite similar to those of the LBGs seen at higher redshifts, and they may indeed be closely related to the LBG population.
Lyman-break galaxies at high redshift. By variation of the filter set, drop-outs can also be discovered at larger wavelengths, thus at accordingly higher redshifts. The object selection at higher z implies an increasingly dominant role of the Lyα forest whose density is a strongly increasing function of redshift (see Sect. 8.​5.​2). This method has been routinely applied with ground-based observations up to z ∼ 4. 5, yielding so-called B-drop-outs. Galaxies at considerably higher redshifts are difficult to access from the ground with this method. One reason for this is that galaxies become increasingly faint with redshift, rendering observations substantially more difficult. Furthermore, one needs to use increasingly redder filter sets. At such large wavelengths the night sky gets significantly brighter, which further hampers the detection of very faint objects. For detecting a galaxy at redshift, say, z = 5. 5 with this method, the Lyα line, now at λ ≈ 7900 Å, is located right in the I-band, so that for an efficient application of the drop-out technique only the I- and z-band filters or NIR-filters are viable, and with those filters the brightness of the night sky is very problematic (see Fig. 9.6 for an example of a drop-out galaxy at very high redshift). Furthermore, candidate very high-redshift galaxies detected as drop-outs are very difficult to verify spectroscopically due to their very faint flux and the fact that most of the diagnostic spectral features are shifted to the near-IR. In spite of this, we will see later that the drop-out method has achieved spectacular results even at redshifts considerably higher than z ∼ 4, where the HST played a central role.

9.1.2 Photometric redshift

Spectral breaks. The Lyman-break technique is a special case of a method for estimating the redshift of galaxies (and QSOs) by multi-color photometry. This technique can be employed due to the spectral breaks at λ = 912 Å and λ = 1216 Å, respectively. Spectra of galaxies also show other characteristic features. As was discussed in detail in Sect. 3.​5, the broad-band energy distribution is basically a superposition of stellar radiation (if we ignore for a moment the presence of dust, which can yield a substantial infrared emission from galaxies). A stellar population of age ≳ 108 yr features a 4000 Å-break because, due to a sudden change in the opacity at this wavelength, the spectra of most stars show such a break at about 4000 Å (see Fig. 3.​33). Hence, the radiation from a stellar population at λ < 4000 Å is less intense than at λ > 4000 Å; this is the case particularly for early-type galaxies, as can be seen in Fig. 3.​36, due to their old stellar population.
Principle of the method. If we assume that the star-formation histories of galaxies are not too diversified, the spectral energy distributions of these galaxies are expected to follow certain patterns. For example, if all galaxies had a single episode of star formation, starting at redshift z f and lasting for a time τ, then the spectra of these galaxies, for a given initial mass function, would be characterized by these two parameters, as well as the total stellar mass formed (see Sect. 3.​5); this latter quantity then yields the amplitude of the spectrum, but does not affect the spectral shape. When measuring the magnitude of these galaxies in n broad-band filters, we can form n − 1 independent colors. Next suppose we form a multi-dimensional color-color diagram, in which every galaxy is represented by a point in this (n − 1)-dimensional color space. Considering only galaxies at the present epoch, all these points will lie on a two-dimensional surface in this multi-dimensional space, instead of being more or less randomly distributed.
Next, instead of plotting z = 0 galaxies, we consider the distribution of galaxies at some higher redshift z < z f. This distribution of points will be different, mainly due to two different effects. First, a given photometric filter corresponds to a different rest-frame spectral range of the galaxy, due to redshift. Second, the ages of the stellar populations are younger at an earlier cosmic epoch, and thus the spectral energy distributions are different. Both of these effects will cause these redshift z galaxies to occupy a different two-dimensional surface in multi-color space. Generalizing this argument further, we see that in this idealized consideration, galaxies will occupy a three-dimensional subspace in (n − 1)-dimensional color space, parametrized by formation redshift z f, time-scale τ and the galaxy’s redshift z. Hence, from the measurement of the broad-band energy distribution of a galaxy, we might expect to be able to determine, or at least estimate, its redshift, as well as other properties such as the age of its stellar population; this is the principle of the method of photometric redshifts.
Of course, the situation is considerably more complicated in reality. Galaxies most likely have a more complicated star-formation history than assumed here, and hence they will not be confined to a two-dimensional surface at fixed redshift, but instead will be spread around this surface. The spectrum of a stellar population also depends on its metallicity, as well as absorption, either by gas and dust in the interstellar medium or hydrogen in intergalactic space (of which the Lyman-break method makes proper use). On the other hand, we have seen in Sect. 3.​6 that the colors of current-day galaxies are remarkably similar, best indicated by the red sequence. Therefore, the method of photometric redshifts may be expected to work, even if more complications are accounted for than in the idealized example considered above.
The method is strongly aided if the galaxies have characteristic spectral features, which shift in wavelength as the redshift is changed. If, for example, the spectrum of a galaxy was a power law in wavelength, then the redshifted spectrum would as well be a power law, with the same spectral slope—if we ignore the different age of the stellar population. Therefore, for such a spectral energy distribution is would be impossible to estimate a redshift. However, if the spectrum shows a clear spectral break, then the location of this break in wavelength depends directly on the redshift, thus yielding a particularly clean diagnostic. In this context the 4000 Å-break and the Lyα-break play a central role, as is illustrated in Fig. 9.7.
Calibration. In order to apply this method, one needs to find the characteristic domains in color space where (most of) the galaxies are situated. This can be done either empirically, using observed energy distributions of galaxies, or by employing population synthesis model. More precisely, a number of standard spectra of galaxies (so-called templates) are used, which are either selected from observed galaxies or computed by population synthesis models. Each of these template spectra can then be redshifted in wavelength. For each template spectrum and any redshift, the expected galaxy colors are determined by integrating the spectral energy distribution, multiplied by the transmission functions of the applied filters, over wavelength [see (A.25)]. This set of colors can then be compared with the observed colors of galaxies, and the set best resembling the observation is taken as an estimate for not only the redshift but also the galaxy type.
A129044_2_En_9_Fig7_HTML.gif
Fig. 9.7
The bottom panel illustrates again the principle of the drop-out method, for a galaxy at z ∼ 3. 2. Whereas the Lyman-α forest absorbs part of the spectral flux between (rest-frame wavelength) 912 and 1216 Å, the flux below 912 Å vanishes almost completely. By using different combinations of filters (top panel), an efficient selection of galaxies at other redshifts is also possible. The example shows a galaxy at z = 1 whose 4000 Å-break is located between the two redder filters. The 4000 Å-break occurs in stellar populations after several 107 yr (see Fig. 3.​33) and is one of the most important features for the method of photometric redshift. Source: K.L. Adelberger 1999, Star Formation and Structure Formation at Redshifts 1 < z < 4, astro-ph/9912153, Fig. 1
/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig8_HTML.jpg
Fig. 9.8
Photometric redshift versus the spectroscopic redshift for galaxies in the HDF-North. Photometric data in four optical and two NIR bands have been used here. We see how accurate photometric redshifts can be—their quality depends on the photometric accuracy in the individual filters, the number of filters used, the redshift and the type of the galaxy, and also on details of the applied analysis method. Source: N. Benítez 2000, Bayesian Photometric Redshift Estimation, ApJ 536, 571, p. 579, Fig. 5. ©AAS. Reproduced with permission
Practical considerations. The advantage of this method is that multi-color photometry is much less time-consuming than spectroscopy of galaxies. Whereas some modern spectrographs allow one to take spectra of ∼ 1000 objects simultaneously, images taken with wide-field cameras of ∼ 1 deg2 on 4-m class telescopes record the fluxes of ∼ 105 galaxies in a one hour exposure. In addition, this method can be extended to much fainter magnitudes than are achievable for spectroscopic redshifts. The disadvantage of the method becomes obvious when an insufficient number of photometric bands are available, since then the photometric redshift estimates can yield a completely wrong z; these are often called catastrophic outliers. One example for the occurrence of extremely wrong redshift estimates is provided by a break in the spectral energy distribution. Depending of whether this break is identified as the Lyman-break or the 4000 Å-break, the resulting redshift estimates will be very different. To break the corresponding degeneracy, a sufficiently large number of filters spread over a broad spectral range must be available to probe the spectral energy distribution over a wide range in wavelengths. As a general rule, the more photometric bands are available and the smaller the uncertainties in the measured magnitudes, the more accurate the estimated redshift. Normally, data from four or five photometric bands are required to obtain useful redshift estimates. In particular, the reliability of the photometric redshift benefits from data over a large wavelength range, so that a combination of several optical and NIR filters is desirable.
The successful application of this method also depends on the type of the galaxies. As we have seen in Sect. 6.​8, early-type galaxies form a relatively well-defined color-magnitude sequence at any redshift, due to their old stellar populations (manifested in clusters of galaxies in form of the red cluster sequence), so that the redshift of this type of galaxy can be estimated very accurately from multi-color information. However, this is only the case if the 4000 Å-break is located in between two of the applied filters. For z ≳ 1 this is no longer the case if only optical filters are used. Other types of galaxies show larger variations in their spectral energy distribution, depending, e.g., on the star formation history, as mentioned before.
Photometric redshifts are particularly useful for statistical purposes, for instance in situations in which the exact redshift of each individual galaxy in a sample is of little relevance. However, by using a sufficient number of optical and NIR filters, quite accurate redshift estimates for individual galaxies are achievable. For example, with eight optical and NIR bands and accurate photometry, a redshift accuracy of Δ z ∼ 0. 03(1 + z) was obtained, as demonstrated in Fig. 9.8 by a comparison of photometric redshifts with redshifts determined spectroscopically for galaxies in the field of the HDF-North. If data in additional photometric bands are available, e.g., using filters of smaller transmission curves (‘intermediate width filters’), the redshift accuracy can be increased even more, e.g., Δ z ∼ 0. 01(1 + z) was obtained using a total of 30 photometric bands.

9.1.3 Other few-band selection techniques

The Lyman-break technique is a special case of the photometric redshift method; it relies on only three photometric bands to select galaxies in a given redshift range, whereas in general, more bands are needed to obtain reliable redshift estimates. There are other cases where a few bands are sufficient for a fairly reliable selection of particular kinds of galaxies, or particular redshift regimes, some of which should be mentioned here.
Selection of 1. 5 ≲ z ≲ 2. 5 galaxies. The success of the Lyman-break method is rooted in the fact that the observed colors of star-forming galaxies in a carefully selected triplet of filters is essentially independent on details of the star-formation history, metallicity and other effects, due to a very strong spectral break. This is illustrated in Fig. 9.4. The same figure also shows that the colors of galaxies with somewhat lower redshift are also very similar; for example, one sees that galaxies with z ∼ 1. 8 all have U nG ∼ 0 and  $$G -\mathcal{R}\sim 0$$ . At that redshift, the Lyα-line is shortward of the U n-band filter, and thus a star-forming galaxy has a rather flat spectrum across all three filters. As the redshift increases above z ∼ 2, the Lyα-line moves into the Un-band filter and thus increases the flux there; however, as we have seen, a large fraction of LBGs have rather low Lyα-flux, thus affecting the color only marginally. For redshifts higher than ∼ 2. 5, the break moves into the Un-band, and the objects redden in U nG and move onto the same sequence where LBGs are selected. Thus, with a single set of three filters (and thus the same optical images), one can select galaxies over the broad range of 1. 5 ≲ z ≲ 3. 5.
A129044_2_En_9_Fig9_HTML.gif
Fig. 9.9
Two color diagram (Bz) vs (zK) for K-band selected galaxies of the K20 survey in the GOODS field. Red solid triangles and circles denote star-forming and passive galaxies, respectively, at z ≥ 1. 4, and blue open squares correspond to additional z ≥ 1. 4 objects as determined from their photometric redshifts. Black solid squares are galaxies with redshift below 1.4, and the green asterisks are stars. Encircles symbols are galaxies detected in X-rays. The various lines delineate regions of photometric selection of z > 1. 4 galaxies—see text. Source: E. Daddi et al. 2004, A New Photometric Technique for the Joint Selection of Star-forming and Passive Galaxies at 1. 4 ≲ z ≲ 2. 5, ApJ 617, 746, p. 749, Fig. 3. ©AAS. Reproduced with permission
BzK selection. While the filter combination used for the Lyman-break galaxies selects star-forming galaxies at high redshift, it misses galaxies with a passive stellar population. One has therefore investigated whether another combination of filters, and thus different colors, may be able to identify high-redshift passive galaxies. Indeed, such a filter set was found; the combination of the B-, z- and K-band filters provides a successful tool to search for galaxies with 1. 4 ≲ z ≲ 2. 5, as illustrated in Fig. 9.9. K-band selected galaxies with 1. 4 ≲ z ≲ 2. 5 occupy specific regions in a Bz versus zK color-color diagram.2 In this redshift range, the 4000 Å-break is located redward of the z-band, thus such galaxies display a fairly red zK color if they are not forming stars at a high rate. The lack of active star formation also causes the Bz color to be rather red, since the B-band probes to the rest-frame UV-region of the spectrum. Such galaxies are located in the upper right corner of the diagram in Fig. 9.9. In case the galaxies in this redshift range are actively forming stars, the 4000 Å-break is weaker, but instead the Bz color is rather blue, so that these galaxies are located in the upper left corner of the diagram. As Fig. 9.9 shows, this selection of high-redshift galaxies is very efficient.
The BzK-selected galaxies with active star formation have redder colors in the rest-frame UV than the Lyman-break galaxies which are selected based on their UV flux, although there is a significant overlap between the two populations in the sense that a substantial fraction of galaxies are found by both methods. However, the most actively star-forming galaxies are missed with the BzK-method since those show little-to-no 4000 Å-break, thus no longer have a sufficiently red zK color, and would lie below the solid line in Fig. 9.9.
/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig10_HTML.jpg
Fig. 9.10
The evolution of (JK) color as a function of redshift. Solid curves show the color for different ages of the stellar population. Dashed and dotted curves correspond to stellar populations with continuous star formation, for different ages and reddening. The dash-dotted curve corresponds to a single age population formed at z = 5. The color is redder than  $$(J - K) = 2.3$$ for the single-age stellar populations at z > 2. 5, and for the one formed at z = 5, this color criterion is satisfied for all z > 2. Source: M. Franx et al. 2003, A Significant Population of Red, Near-Infrared-selected High-Redshift Galaxies, ApJ 587, L79, p. L80, Fig. 1. ©AAS. Reproduced with permission
Distant red galaxies. Another method to select high-redshift passive galaxies is based on their rest-frame optical colors. From local galaxies we know that the 4000 Å-break is the most prominent feature in the spectral energy distribution of stellar populations with no or little star formation. At redshifts 2 ≲ z ≲ 4, this break is located between the observed J- and K-band filters; hence we expect that passive galaxies are red in their JK color. As Fig. 9.10 shows, JK ≳ 2. 3 as soon as the redshift increases beyond z ≳ 2. Perhaps surprisingly, this is true even if the stellar population is as young as 0. 25 Gyr, for which the redshift of the transition to JK > 2. 3 occurs at only slighter larger redshift. Furthermore, this color selection is able to find also galaxies with ongoing star formation, provided they also have an old stellar population; this is due to the fact that much of the star formation is accompanied by substantial dust obscuration. At redshift z = 2, the J-band corresponds to the rest-frame B-band, which is substantially affected by extinction, leading to a red JK color. High-redshift galaxies selected according to JK > 2. 3 are called distant red galaxies (DRGs). The fact that there is very little overlap in the galaxy population selected according to their UV-properties and the DRG population immediately shows the necessity to apply several very different selection criteria for high-redshift galaxies to obtain a complete census of their population.
Narrow-band selection. We mentioned the method of narrow-band selection before. If a source has a strong emission line, and if the observed wavelength of the emission line matches the spectral response of a narrow-band filter, then the ratio of fluxes obtained in this narrow-band image compared to a broad-band image would be much larger than for other sources without a strong emission line at the corresponding wavelength.
After a substantial population of high-redshift galaxies were found with the Lyman Break technique, it became known that about 60 % of these galaxies show very strong Lyα emission lines. It was then possible to design narrow-band filters that were particularly tuned to detect objects with strong Lyα emission lines at a particular redshift. Several thousand Lyα emitters (LAEs) were detected with this method, extending up to redshift z ∼ 7. These galaxies are on average considerably fainter than LBGs, and therefore allow one to probe the fainter end of the luminosity function of star-forming galaxies. Their faintness, on the other hand, make more detailed spectroscopic studies very challenging, and thus the relation of these Lyα emitters to the other galaxy populations at similar redshifts is not easy to determine.
Furthermore, candidate objects detected in narrow-band images require spectroscopic follow-up, since there are many possible contaminants that may enter the selection. Galaxies, and in particular AGNs, at lower redshifts can display strong emission lines of other atomic transitions and need to be ruled out with a spectrum. Due to the cumulative effect of the Lyα forest, a high-redshift (z ≳ 4) Lyα emitter should show essentially no flux at shorter wavelengths, and so some of the Lyα emission-line candidates can be rejected if continuum flux bluewards of the narrow band is detected.

9.2 Deep views of the Universe

Very distant objects in the Universe are expected to be exceedingly faint. Therefore, in order to find the most distant, or earliest, objects in the Universe, very deep images of the sky are needed to have a chance to detect them.
In order to get further out into the Universe, astronomers use their most sensitive instruments to obtain extremely deep sky images. The Hubble Deep Field, already discussed briefly in Sect. 1.​3.​3, is perhaps the best-known example for this. As will be discussed below, further instrumental developments have led to even deeper observations with the HST. Deep fields are taken also with ground-based optical and near-IR telescopes. Although the sensitivity limit from the ground is affected by the atmosphere, in particular at longer wavelengths, this drawback is partly compensated by the larger field-of-view that many ground-based instruments offer, compared to the relatively small field-of-view of the HST. Dep field observations are conducted also at other wavelengths, preferentially in the same sky areas as the deep optical fields, to enable cross-identification and thus provide additional information on the detected sources. As we shall see, the availability of such deep fields has allowed us to take a first look at the first 10 % of the Universe’s life.

9.2.1 Hubble Deep Fields

The Hubble Deep Field North. In 1995, an unprecedented observing program was conducted with the HST. A deep image in four filters (U300, B450, V606, and I814) was observed with the Wide Field/Planetary Camera 2 (WFPC2) on-board HST, covering a field of ∼ 5. 3 arcmin2, with a total exposure time of about 10 days. This resulted in the deepest sky image of that time, displayed in Fig. 1.​37. The observed field was carefully selected such that it did not contain any bright sources. Furthermore, the position of the field was chosen such that the HST was able to continually point into this direction, a criterion excluding all but two relatively small regions on the sky, due to the low HST orbit around the Earth. Another special feature of this program was that the data became public right after reduction, less than a month after the final exposures had been taken. Astronomers worldwide immediately had the opportunity to scientifically exploit these data and to compare them with data at other frequency ranges or to perform their own follow-up observations. Such a rapid and wide release was uncommon at that time, but is now seen more frequently. Rarely has a single data set inspired and motivated a large community of astronomers as much as the Hubble Deep Field (HDF) did (after another HDF was observed in the Southern sky—see below—the original HDF was called HDF North, or HDFN).
Follow-up observations of the HDF were made in nearly all accessible wavelength ranges, so that it became the best-observed region of the extragalactic sky. Within a few years, more than ∼ 3000 galaxies, 6 X-ray sources, 16 radio sources, and fewer than 20 Galactic stars were detected in the HDF, and redshifts were determined spectroscopically for more than 150 galaxies in this field, with about 30 at z > 2. Never before could galaxy counts be conducted to magnitudes as faint as it became possible in the HDF (see Fig. 9.11); several hundred galaxies per square arcminute could be photometrically analyzed in this field.
Detailed spectroscopic follow-up observations were conducted by several groups, through which the HDF became, among other things, a calibration field for photometric redshifts (see, for instance, Fig. 9.8). Most galaxies in the HDF are far too faint to be analyzed spectroscopically, so that one often has to rely on photometric redshifts.
/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig11_HTML.jpg
Fig. 9.11
Galaxy counts from the HDF and other surveys in four photometric bands: U, B, I, and K. Solid symbols are from the HDF, open symbols from various ground-based observations. The curves represent predictions from models in which the spectral energy distribution of the galaxies does not evolve—the counts lie significantly above these so-called non-evolution models: clearly, the galaxy population must be evolving. Note that the counts in the different color filters are shifted by a factor 10 each, simply for display purposes. Source: H. Ferguson et al. 2000, The Hubble Deep Fields, ARA&A 38, 667, Fig. 4. Reprinted, with permission, from the Annual Review of Astronomy & Astrophysics, Volume 38 ©2000 by Annual Reviews www.​annualreviews.​org
HDFS and the Hubble Ultra Deep Field. Later, in 1998, a second HDF was observed, this time in the southern sky. In contrast to the HDFN, which had been chosen to be as empty as possible, the HDFS contains a QSO. Its absorption line spectrum can be compared with the galaxies found in the HDFS, by which one hopes to obtain information on the relation between QSO absorption lines and galaxies. In addition to the WFPC2 camera, the HDFS was simultaneously observed with the cameras STIS (51″ × 51″ field-of-view, where the CLEAR ‘filter’ was used, which has a very broad spectral sensitivity; in total, STIS is considerably more sensitive than WFPC2) and NICMOS (a NIR camera with a maximum field-of-view of 51″ × 51″) which had both been installed in the meantime. Nevertheless, the overall impact of the HDFS was smaller than that of the HDFN; one reason for this may be that the requirement of the presence of a QSO, combined with the need for a field in the continuous viewing zone of HST, led to a field close to several very bright Galactic stars. This circumstance makes photometric observations from the ground very difficult, e.g., due to stray-light.
One of the immediate results from the HDF was the finding that the morphology of faint galaxies is quite different from those in the nearby Universe. Locally, most luminous galaxies fit into the morphological Hubble sequence of galaxies. This ceases to be the case for high-redshift galaxies. In fact, galaxies at z ∼ 2 are much more compact than local luminous galaxies, they show irregular light distributions and do not resemble any of the Hubble sequence morphologies. By redshifts z ∼ 1, the Hubble sequence seems to have been partly established.
In 2002, an additional camera was installed on-board HST. The Advanced Camera for Surveys (ACS) has, with its side length of 3.​​4, a field-of-view about twice as large as WFPC2, and with half the pixel size (0.​​05) it better matches the diffraction-limited angular resolution of HST. Therefore, ACS is a substantially more powerful camera than WFPC2 and is, in particular, best suited for surveys. With the Hubble Ultra Deep Field (HUDF), the deepest image of the sky was observed and published in 2004 (see Fig. 9.12). The HUDF is, in all four filters, deeper by about one magnitude than the HDF, reaching a magnitude limit of m AB ≈ 29. The depth of the ACS images in combination with the relatively red filters that are available provides us with an opportunity to identify drop-out candidates out to redshifts z ∼ 6; quite a number of such candidates have already been verified spectroscopically.
Lyman-break galaxies at z ∼ 6 seem to have stellar populations with masses and lifetimes comparable to those at z ∼ 3. This implies that at a time when the Universe was 1 Gyr old, a stellar population with mass ∼ 3 × 1010 M and age of a few hundred million years (as indicated by the observed 4000 Å break) was already in place. This, together with the apparently high metallicity of these sources, is thus an indication of how quickly the early Universe evolved. The z ∼ 6 galaxies are very compact, with half-light radii of ∼ 1 kpc, and thus differ substantially from the galaxy population known in the lower-redshift Universe.
Half of the HUDF was also imaged by the near-IR NICMOS camera onboard HST, but only after the installment
/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig12_HTML.jpg
Fig. 9.12
The Hubble Ultra Deep Field, a field of ∼ 3.​​4 × 3.​​4 observed by the ACS camera. The limiting magnitude up to which sources are detected in this image is about one magnitude fainter than in the HDF. More than 10 000 galaxies are visible in the image, many of them at redshifts z ≥ 5. Credit: NASA, ESA, S. Beckwith/Space Telescope Science Institute, and the HUDF Team
of the Wide-Field Camera 3 (WFC3) on HST, with a near-IR channel and a much larger field-of-view and much better sensitivity than NICMOS, could the HUDF be imaged to comparable sensitivity levels in the NIR as in the optical. WFC3 mapped the HUDF in three NIR bands, Y, J and H, down to a limiting magnitude of m AB ≈ 28. 5. These long wavelengths allowed the systematic search for Lyman-break galaxies at redshifts beyond 6, as will be discussed below.
/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig13_HTML.jpg
Fig. 9.13
The Hubble eXtreme Deep Field (XDF) covers an area of 2.​​3 × 2′, centered on the HUDF, and was composed with HST observations spread over many years. The total exposure time amounts to about 2 million seconds, or 22 days. This color composite was made from data in eight different optical and NIR bands, taken with the ACS and WFC3/IR instruments. Credit: NASA, ESA, G. Illingworth, D. Magee, and P. Oesch (University of California, Santa Cruz), R. Bouwens (Leiden University), and the HUDF09 Team
In September 2012, the deepest view of the Universe ever taken was released: the eXtreme Deep Field (XDF), shown in Fig. 9.13. It covers about 4. 7 arcmin2 in nine optical and NIR filters, reaching a limiting magnitude of m AB ≈ 30.
Further deep-field projects with HST: GOODS, GEMS, COSMOS. The great scientific harvest from the deep HST images, particularly in combination with data from other telescopes and the readiness to make such data available to the scientific community for multi-frequency analysis, provided the motivation for additional HST surveys. The GOODS (Great Observatories Origins Deep Surveys) project is a joint observational campaign of several observatories, centering on two fields of ∼ 16′ × 10′ size each that have been observed by the ACS camera at several epochs between 2003 and 2005. One of these two regions (GOODS-North) contains the HDFN, the other a field that became known as the Chandra Deep Field South (CDFS), also containing the HUDF. The Chandra satellite observed both GOODS fields with a total exposure time of ∼ 2 × 106 s and ∼ 4 × 106 s, corresponding to about 24 and 48 days, respectively. Also, the Spitzer observatory took long exposures of these two fields, and several ground-based observatories are involved in this survey, for instance by contributing an ultra-deep wide-field image ( ∼ 30′ × 30′) centered on the Chandra Deep Field South and NIR images in the K-band. The data themselves and the data products (like object catalogs, color information, etc.) are all publicly available and have led to a large number of scientific results.
The multi-wavelength approach by GOODS yields an unprecedented view of the high-redshift Universe. Although these studies and scientific analysis are ongoing (at the time of writing), quite a large number of very high-redshift (z ≳ 5–6) galaxies were discovered and studied: a sample of more than 500 I-band drop-out candidates was obtained from deep ACS/HST images.
Even larger surveys were conducted with the HST, including the Galaxy Evolution from Morphology and SEDs (GEMS) survey, covering a field of 30′ × 30′ centered on the CDFS mapped in two filters. For this field, full coverage with a 17-band (5 broad bands, and 12 intermediate width bands) optical imaging survey (COMBO17) is available. The largest contiguous field imaged with the angular resolution of HST is the ∼ 2 deg2 COSMOS survey. This sky area was also imaged with other space-based (Spitzer, GALEX, XMM-Newton, Chandra) and ground-based (Subaru, VLA, VLT, UKIRT and others) observatories. Its large field enables the study of the large-scale galaxy and AGN distribution at high redshifts. The broad wavelength coverage from the radio to the X-ray regime, renders the COSMOS field a treasure for observational cosmology for years to come. As discussed in Sect. 8.​4, the COSMOS field was used for a detailed cosmic shear analysis, where the broad wavelength coverage helped enormously to determine accurate photometric redshifts of the source galaxies.

9.2.2 Deep fields in other wavebands

Deep fields have been observed in many other frequency ranges, and as mentioned before, they are often taken on the same sky areas to allow for multi-frequency studies of the sources. We mention here the Chandra Deep Fields, one in the North (CDFN) in the direction of the HDFN, the other in the South (CDFS, see Fig. 9.14) in the direction of the HUDF, with a total exposure time of two and four million seconds, respectively.
/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig14_HTML.jpg
Fig. 9.14
The Chandra Deep Field South (CDFS), an X-ray image of a ∼ 450 arcmin2 field with a total exposure time of 4 × 106 s. This is the deepest image ever taken of the X-ray sky. The right panel shows a color composite image taken with HST. Credit: X-ray: NASA/CXC/U.Hawaii/E.Treister et al.; optical: NASA/STScI/S.Beckwith et al.)
/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig15_HTML.jpg
Fig. 9.15
Cumulative compact source counts N( > S) as obtained from the CDFS (see Fig. 9.14), in the soft (a) and hard (b) energy bands, split according to source populations: AGNs (blue), galaxies (red) and Galactic stars (green). Source: B.D. Lehmer et al. 2012, The 4 Ms Chandra Deep Field-South Number Counts Apportioned by Source Class: Pervasive Active Galactic Nuclei and the Ascent of Normal Galaxies, ApJ 752:46, p. 7, Fig. 5. ©AAS. Reproduced with permission
/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig16_HTML.jpg
Fig. 9.16
HST images of the lensing cluster Abell 68 magnifying many faint background galaxies. Several z > 5 candidates are found in this field, at least one of them spectroscopically verified. Credit: NASA, ESA, and the Hubble Heritage/ESA-Hubble Collaboration, Acknowledgment: N. Rose
A129044_2_En_9_Fig17_HTML.gif
Fig. 9.17
The image on the left was taken by the Hubble Space Telescope. It shows the cluster of galaxies MS 1512+36, which has a redshift of z = 0. 37. To the right, and slightly above the central cluster galaxy, an extended and apparently very blue object is seen, marked by an arrow. This source is not physically associated with the cluster but is a background galaxy at a redshift of z = 2. 72. With this HST image it was proved that this galaxy is strongly lensed by the cluster and, by means of this, magnified by a factor of ∼ 30. Due to the magnification, this Lyman-break galaxy is the brightest normal galaxy at redshift z ∼ 3, a fact that can be profitably used for a detailed spectroscopic analysis. On the right, a small section from a high-resolution VLT spectrum of this galaxy is shown. The Lyα transition of the galaxy is located at λ = 4530 Å, visible as a broad absorption line. Absorption lines at shorter wavelengths originate from the Lyα-forest along the line-of-sight (indicated by short vertical lines) or by metal lines from the galaxy itself (indicated by arrows). Credit: Left: S. Seitz, HST; corresponding research article: S. Seitz et al. 1998, The z = 2. 72 galaxy cB58: a gravitational fold arc lensed by the cluster MS1512+36, MNRAS 298, 945. Right: European Southern Observatory
With the extremely faint limiting flux of the CDFS, the source density on the sky reaches more than 104 deg−2, which is comparable to the source density seen in rather shallow optical imaging, like the SDSS images. However, the mean redshift of the X-ray sources is considerably higher than that of the optical sources, due to the large fraction of AGNs. As we can see from Fig. 9.15, the X-ray population of compact sources in ultra-deep X-images is composed of three components. By far dominating is the population of AGNs, whose number counts exhibit the shape of a broken power law—steep at the bright end, flatter at the low-flux end. For high fluxes, they follow approximately the Euclidean slope,  $$\mathrm{d}N/\mathrm{d}S \propto S^{-2.5}$$ , whereas at fluxes below the breakpoint at  $$S_{\mathrm{b}} \sim 6 \times 10^{-15}\,\mathrm{erg\,cm^{-2}\,s^{-1}}$$ , one finds  $$\mathrm{d}N/\mathrm{d}S \propto S^{-1.5}$$ for the soft band, and even slightly flatter in the hard band. Going to fainter X-ray fluxes, the fraction of obscured AGNs at high redshifts increases.
Approaching the fainter flux levels, the population of normal galaxies becomes increasingly important. At the limiting flux of the CDFS, they constitute almost 50 % of the source population. They are mainly late-type star-forming galaxies, with the X-ray emission mostly due to X-ray binaries. Early-type galaxies contribute just a small fraction of ∼ 10 % to the galaxy counts, where their X-ray emission is a due to a combination of X-ray binaries and hot gas.

9.2.3 Natural telescopes

Galaxies at high redshift are faint and therefore difficult to observe spectroscopically. For this reason, the brightest galaxies are preferentially selected (for detailed examination), i.e., basically those which are the most luminous ones at a particular z—resulting in undesired, but hardly avoidable selection effects. For example, those Lyman-break galaxies at z ∼ 3 for which the redshift is verified spectroscopically are typically located at the bright end of the LBG luminosity function. The sensitivity of our telescopes is insufficient in most cases to spectroscopically analyze a rather more typical galaxy at z ∼ 3.
The magnification by gravitational lenses can substantially boost the apparent magnitude of sources; gravitational lenses can thus act as natural (and inexpensive!) telescopes. The most prominent examples are the arcs in clusters of galaxies: many of them have a very high redshift, are magnified by a factor ≳ 5, and hence are brighter by about ≳ 1. 5 mag than they would be without the lens effect (see Fig. 9.16).3 In addition to boosting the observable fluxes, the gravitational lens effect yields a spatial magnification of the sources: the sources appear larger than they really are, thus increasing the effective angular resolution with which they can be observed.
Lyman-break galaxies at z ∼ 3. An extreme first example of this effect is represented by the galaxy cB58 at z = 2. 72, which is displayed in Fig. 9.17. It was discovered in the background of a galaxy cluster and is magnified by a factor ∼ 30. Hence, it appears brighter than a typical Lyman-break galaxy by more than three magnitudes. For this reason, one of the most detailed spectra of all galaxies at z ∼ 3 was taken of this particular source. Several more such examples were found subsequently, including the so-called Cosmic Eye, a Lyman-break galaxy at z = 3. 07, shown in Fig. 9.18. The typical magnification of these lensed LBGs is μ ∼ 30.
Further examples. Furthermore, at least two highly magnified Lyα emitters (LAEs) at z ≈ 5 were found. Whereas the study of LAEs is usually hampered by the faint continuum, the strong magnification (μ > 10 in these two systems) enables an investigation of the underlying stellar population from broad-band photometry. One of these two systems has an estimated stellar mass of ≲ 108 M , one of the lowest mass systems found so far at high redshifts. Thanks to the increased effective angular resolution provided by gravitational lensing magnification, a spatially resolved view of the star-forming regions in a z = 2. 33 sub-millimeter galaxy was obtained. The observations are compatible with the picture that star formation in this object occurs in the cores of giant molecular clouds, as in the local Universe, but these regions are ∼ 100 times larger than local star-forming sites.
/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig18_HTML.jpg
Fig. 9.18
The Cosmic Eye is a Lyman-break galaxy at z = 3. 07, gravitationally lensed by an early-type galaxy at z = 0. 73, which itself is located in the background of a massive galaxy cluster at z = 0. 33. The angular extent of the arc systems is ∼ 3″; this large image splitting is due to the combined lensing effect of the main lens galaxy and the foreground cluster, which magnify the source galaxy by a factor μ ∼ 30. Hence, this image is almost 4 magnitudes brighter than the unlensed source. Credit: D.P. Stark, HST
A129044_2_En_9_Fig19_HTML.gif
Fig. 9.19
A section of the galaxy cluster Abell 2218 (z = 0. 175), observed with the HST in four different filters. This region was selected because the magnification by the gravitational lens effect for sources at high redshift is expected to be very large here. This fact has been established by a detailed mass model of this cluster which could be constructed from the geometrical constraints provided by the numerous arcs and multiple images (Fig. 6.​51). The red lines denote the critical curves of this lens for source redshifts of z = 5, 6.5, and 7. A double image of an extended source is clearly visible in the NIR image (on the right); this double image was not detected at shorter wavelengths—the expected position is marked by two ellipses in the two images on the left. The direction of the local shear, i.e., of the expected image distortion, is plotted in the second image from the right; the observed elongation of the two images a and b is compatible with the shear field from the lens model. Together with the photometry of these two images, a redshift between z = 6. 8 and z = 7 was derived for the source of this double image. Source: J.-P. Kneib et al. 2004, A Probable z ∼ 7 Galaxy Strongly Lensed by the Rich Cluster A2218: Exploring the Dark Ages, ApJ 607, 697, p. 698, Fig. 1. ©AAS. Reproduced with permission
Apparently extreme sources. One can argue that there is a high probability that the flux of the apparently most luminous sources from a particular source population is magnified by lensing. The apparently most luminous IRAS galaxy, F10214+47, is magnified by a factor ∼ 50 by the lens effect of a foreground galaxy, where the exact value of the magnification depends on the wavelength, since the intrinsic structure and size of the source is wavelength-dependent—hence the magnification is differential. Other examples are the QSOs B1422+231 and APM 08279+5255, which are among the brightest QSOs despite their high redshifts; hence, they belong to the apparently most luminous sources in the Universe. In both cases, multiple images of the QSOs were discovered, verifying the action of the lens effect. Their magnification, and therefore their brightness, renders these sources preferred objects for QSO absorption line spectroscopy (see Fig. 5.​55). The Lyman-break galaxy cB58 mentioned previously is another example, and we will see below that the most luminous sub-millimeter sources are gravitationally lensed.
Employing natural telescopes. Most of the examples of highly magnified sources mentioned before were found serendipitously. However, the magnification effect can also be utilized deliberately, by searching for high-redshift sources in regions which are known to produce strong magnification effects, i.e., fields around clusters of galaxies: for a massive cluster, one knows that distant sources located behind the cluster center are substantially magnified. It is therefore not surprising that some of the most distant galaxies known have been detected in systematic searches for drop-out galaxies near the centers of massive clusters. One example of this is shown in Fig. 9.19, where a galaxy at z ∼ 7 is doubly imaged by the cluster Abell 2218 (see Fig. 6.​51), and by means of this it is magnified by a factor ∼ 25.
The CLASH and HLS surveys. In order to fully exploit the power of natural telescopes, a large treasury program was carried out with HST, imaging 25 massive clusters with 16 filters of the ACS and WFC3/IR instruments. This CLASH (Cluster Lensing And Supernova survey with Hubble) survey is designed to obtain accurate spectral energy distributions of very faint galaxies in the cluster fields. Furthermore, the data will allow high-quality strong and weak lensing analysis of these clusters, and many results are already available at the time of writing. The observing strategy with multiple visits per cluster field also enables the detection of supernovae, either of cluster galaxies, or gravitationally lensed background supernovae.
/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig20_HTML.jpg
Fig. 9.20
Left panel: Similar to Fig. 9.4, evolutionary tracks of stellar populations in a near-IR color-color diagram are shown, for different galaxy types. The two blue tracks correspond to starburst galaxies, and they differ in the assumed slope β of the continuum UV spectrum. The corresponding redshifts are written near the tracks. Black points show the near-IR colors of galaxies detected at optical wavelengths in the HUDF. The red squares with error bars correspond to five galaxies which are clearly detected in at least two near-IR bands, but with no detection in any optical filter—see the right panel, where cut-outs around these sources in the HUDF optical and WFC3/IR images are shown. These were selected by the color criteria indicated by the grey region. Source: R.J. Bouwens et al. 2010, Discovery of z ∼ 8 Galaxies in the Hubble Ultra Deep Field from Ultra-Deep WFC3/IR Observations, ApJ 709, L133, p. L135, Figs. 1, 2, 3. ©AAS. Reproduced with permission
/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig21_HTML.jpg
Fig. 9.21
Spectrum of the galaxy UDFy-38135539, for which multiband images are shown in Fig. 9.20, obtained with integral field spectrograph SINFONI at the VLT. The grey band at the bottom shows the shape of the night-sky spectrum. The total integration time for this spectrum was 15 h. The Ly α line marked at λ = 11 616 Å is broader than the instrumental resolution, and is detected with ∼ 6σ significance. Source: M.D. Lehnert et al. 2010, Spectroscopic confirmation of a galaxy at redshift z=8.6, Nature 467, 940, Fig. 1. Reprinted by permission of Macmillan Publishers Ltd: Nature, ©2010
In a similar spirit, the Herschel Lensing Survey (HLS) targeted 44 massive clusters, in order to find magnified far-IR sources; some of the results from this survey will be discussed in Sect. 9.3.3 below.

9.2.4 Towards the dark ages

We have seen that in order to apply the Lyman-break technique, one needs images in three filters, one on the blue side of the Lyman break, and two on the red side. These two redder filters are required to demonstrate that these galaxies have a rather flat spectrum for wavelengths longer than the break, i.e., that they are indeed actively star forming, to minimize the possibility that the source is simply a very red one and drops out of the shorter wavelength filter just for this reason. Therefore, the application of the Lyman-break method on deep optical images is restricted to redshifts of z ≲ 5. In order to move towards higher redshifts, images in the near-IR are required.
Combining the images of the HUDF with deep NIR imaging from NICMOS and with ground-based telescopes, the redshift z ∼ 6 barrier could be overcome, by searching for i-band dropouts. Many of these were found, and several of the brighter ones were spectroscopically confirmed. Furthermore, ultra-deep optical and NIR imaging from the ground, covering larger sky areas than possible with the HST, have brought large harvest in selecting z ∼ 6 galaxies using the Lyman-break technique. Currently, more than 30 galaxies are known with spectroscopic redshifts between 6 and 6.5.
/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig22_HTML.jpg
Fig. 9.22
The galaxy cluster MACS J0647+7015 (z = 0. 591) as observed by the HST in the framework of the CLASH survey. The image is a multi-band composite, based on images in 13 different bands. The three small panels on the left of the top image are zooms of a multiply imaged source. The bottom panel shows the critical curves of the lens model for this cluster, for sources at different redshifts of z = 2. 0 (cyan), z = 3. 5 (green) and z = 11. 0 (red). Several strongly lensed images are identified and labeled, where the number of the label identify the same source, and the lower-case letter are assigned to the multiple images. From the lens model of the cluster, which yields an estimate of the magnification μ ≈ 7 for the brightest image, and the multi-band photometry, a likely redshift of z ∼ 10. 7 is estimated for the source. Top: Credit: NASA, ESA, M. Postman and D. Coe (STScI), and the CLASH Team. Bottom: Source: D. Coe et al. 2013, CLASH: Three Strongly Lensed Images of a Candidate z ≈ 11 Galaxy, ApJ 762, 32, p 3, Fig. 1. ©AAS. Reproduced with permission
As we saw before, QSOs at these redshifts have essentially no flux shortward of the Lyα emission line, i.e., the intergalactic medium at z ∼ 6 is sufficiently neutral to absorb almost all UV-photons. This could mean that we are approaching the redshift of reionization of the Universe, and it is not immediately obvious that one may expect a substantial population of higher-redshift objects. On the other hand, the photons needed for reionizing the Universe must come from the first galaxies formed, and these must occur at even higher redshifts. In addition, since the first results from WMAP, we have known that the redshift of reionization is substantially higher than z = 6, with the Planck CMB results estimating this redshift to be at z ∼ 10. Thus, searching for higher-redshift sources is a highly valuable scientific aim, in order to understand the early stages of the evolution of galaxies.
Once at z ∼ 7, the Lyα line is at λ ∼ 1 μm, and hence one needs at least two NIR images, plus a very deep optical image in the z-band, in order to find LBG candidates at these redshifts. As discussed above, the latest camera onboard HST, the WFC3, has a sensitive near-IR channel, and its mapping of the HUDF provided an excellent data set for this purpose. Together with less deep, but larger-area imaging by WFC3 over the GOODS field, some 100 candidates at z ∼ 7, and more than 50 at z ∼ 8 have been found (see Fig. 9.20 for several z ∼ 8 candidates). Spectroscopic verification of these sources, however, is very challenging, since now very sensitive near-IR spectroscopy is needed.
For one of the high-z candidates in Fig. 9.20, the detection of a spectral line in a deep NIR spectrum was claimed—see Fig. 9.21—yielding a redshift of z = 8. 55. Whereas the redshift is based on a single emission line, the identification of this line as Lyα is considered to be secure, since the HST photometry shown in Fig. 9.20 rules out all other redshift, except a low-probability alternative at z ∼ 2. In this case, the emission line would be [Oii], which is a doublet. This doublet would have been clearly resolved in the spectrum and can safely be ruled out. Thus, if the emission line in Fig. 9.21 is real, this galaxy is the highest-redshift spectroscopically confirmed galaxy yet found. However, reobservation of the same source with two other NIR spectrographs failed to reproduce the emission line. Whether or not this galaxy is at z = 8. 55 thus remains an open issue, but this example impressively illustrates the difficulty of securing redshifts for z ≳ 7 galaxies. We also need to keep in mind that many LBGs at lower redshift do not show a Lyα emission line; no spectroscopic redshift of analogs of such sources at high redshifts can be obtained with current telescopes.
Employing natural telescopes, the hunt for even higher-redshift sources becomes promising. The cluster MACS J0647+7015 shown in Fig. 9.22 is a target of the CLASH survey, and thus has been imaged with HST in many different filters. One of the multiply-imaged sources in this cluster is a J-band drop-out; its spectral energy distribution puts it at a redshift z ∼ 11. Such a high redshift is also supported by the lensing geometry and the location of the three images. Whereas no spectroscopic confirmation of this high redshift is available up to now, the next generation space telescope JWST (see Chap. 11) will probably be able to confirm (or not) this redshift estimate.
On the other hand, the narrow-band selection of high-z sources is more promising in this respect. Although there are many contaminating sources—emission line objects at lower redshifts—a clear detection of a source through the narrow-band technique at least yields a good indication that the source has an emission line at that wavelength, and thus spectroscopy is promising.
The wavelengths of the narrow-band filters are best chosen as to minimize the sky brightness. This then defines preferred spectral windows, which in turn define the redshifts of the Lyα emitting galaxies that can be detected with this technique. By far the most productive telescope in this respect is Subaru, due to its unique combination of aperture and field-of-view of its Suprime-Cam camera. Specifically designed narrow-band filters for the highest redshifts target LAEs at z ∼ 5. 7, 6.6, 7.0 and 7.3. Many hundreds of LAEs were found with this techniques, including several dozens spectroscopically confirmed at redshift z ≳ 6. 6. The redshift record holder are galaxies at z = 6. 96 and z = 7. 213.
A129044_2_En_9_Fig23_HTML.gif
Fig. 9.23
Two high-redshift galaxy candidates, selected by the Lyman-break technique from deep HST fields. The data points with error bars are flux measurements from the HST data, complemented by MIR data from Spitzer. The small arrows indicated upper limits to the flux in these bands. The inserts show the figure-of-merit function χ 2 as a function of redshift. The thick blue line in each panel shows the best fitting spectral energy distribution for a high-redshift galaxy, yielding z ∼ 6. 65 and z = 6. 96 for the upper and lower galaxy, respectively, corresponding to the minimum of the χ 2 function. A second spectral fit can be found when assuming a lower redshift, shown by the dotted red curves. In the upper panel, this fit is indeed almost acceptable, though the high-z fit is considerably better—the corresponding χ 2-values as a function of redshift is shown in the inset, and rules out the low-redshift solution with very high confidence. The lower-redshift fit shown in the bottom panel can be rejected fully; this z ∼ 7 estimate seems to be very robust. Source: R.J. McLure et al. 2011, A robust sample of galaxies at redshifts 6.0 < z < 8.7: stellar populations, star formation rates and stellar masses, MNRAS 418, 2074, Fig. 2. Reproduced by permission of Oxford University Press on behalf of the Royal Astronomical Society
/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig24_HTML.jpg
Fig. 9.24
The spectrum of the QSO ULAS J1120+0641 at z = 7. 085 is shown in black, superposed on a composite spectrum of lower-redshift QSOs. Several emission lines redwards of the Lyman-α line are marked. The two spectra are in very good agreement, except of course for rest wavelengths smaller than the Lyman-α line, due to exceedingly strong absorption by the Lyα forest. However, the Civ line of ULAS J1120+0641 seems to be significantly blueshifted, relative to the composite spectrum. Source: D.J. Mortlock et al. 2011, A luminous quasar at a redshift of z = 7.085, Nature 474, 616, Fig. 1. Reprinted by permission of Macmillan Publishers Ltd: Nature, ©2011
In many cases photometric redshifts may be the only method for obtaining redshift information, until more powerful telescopes come into operation. As mentioned before, photometric redshifts are the more reliable the more bands are available, and the better the photometric accuracy is. Deep fields like the GOODS field are therefore best suited for photo-z studies of high-redshift galaxies, owing to the broad wavelength coverage and their depth. In Fig. 9.23, two galaxies selected by the Lyman-break technique are shown, with optical, NIR and MIR observations available from HST and Spitzer. In both cases, the best-fitting high-redshift and low-redshift spectral energy distributions are shown, and in both cases, the low-redshift fit is very much worse than the high-z one, so that a low redshift can be ruled out with very high confidence in both cases. This photo-z technique has yielded substantial samples of z ≲ 8 galaxies, which form a rather robust base for statistical studies of the high-z galaxy population.
The highest redshift QSO. Whereas over most of the past ∼ 50 years QSOs were the redshift record holders, this is currently no longer the case: independent of whether the galaxy shown in Fig. 9.21 is indeed at z = 8. 55, the photo-z of several galaxies are sufficiently robust to place them at z > 7. Concerning QSOs, the SDSS has discovered several z ∼ 6 objects, with the highest redshift one at z = 6. 42 (see Fig. 9.1). This was for many years the record holder; only recently, a QSO with z > 7 was found, whose spectrum is shown in Fig. 9.24. It was found with a color selection, based on optical and near-IR photometry. The spectrum clearly confirms its high redshift of z = 7. 085, based on several emission lines. Its high luminosity of ∼ 6 × 1013 L implies a very massive black hole with mass M ∼ 2 × 109 M , as estimated from the line width of the Mgii emission line in combination with the luminosity. This large SMBH mass had to be assembled within the first 800 million years of the Universe. This strengthens the constraints on rapid black hole formation, relative to the previously discovered QSOs with redshifts z ≲ 6. 4.4

9.3 New types of galaxies

The Lyman-break galaxies discussed above are not the only galaxies that are expected to exist at high redshifts. We have argued that LBGs are galaxies with strong star formation. Moreover, the UV radiation from their newly-born hot stars must be able to escape from the galaxies. From observations in the local Universe we know, however, that a large fraction of star formation is hidden from our direct view when the star formation region is enveloped by dust. The latter is heated by absorbing the UV radiation, and re-emits this energy in the form of thermal radiation in the FIR domain of the spectrum. At high redshifts such galaxies would certainly not be detected by the Lyman-break method.
Instrumental developments opened up new wavelength regimes which yield access to other types of galaxies. Two of these will be described in more detail here: EROs (Extremely Red Objects) and sub-millimeter sources. But before we discuss these objects, we will first investigate starburst galaxies in the relatively local Universe.
/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig25_HTML.jpg
Fig. 9.25
The Antennae galaxies. On the left, the ‘true’ optical colors are shown, whereas in the right-hand image the reddish color shows Hα emission. This pair of merging galaxies (also see Figs. 1.​16 and 3.​6 for other examples of merging galaxies) is forming an enormous number of young stars. Both the UV emission (bluish in the left image) and the Hα radiation (reddish in the right image) are considered indicators of star formation. The individual knots of bright emission are not single stars but star clusters with typically 105 M ; however, it is also possible to resolve individual stars (red and blue supergiants) in these galaxies. Source: B.C. Whitmore et al. 1999, The Luminosity Function of Young Star Clusters in “the Antennae” Galaxies (NGC 4038-4039), AJ 118, 1551, p. 1556, 1557, Figs. 3, 4. ©AAS. Reproduced with permission

9.3.1 Starburst galaxies

One class of galaxies, the so-called starburst galaxies, is characterized by a strongly enhanced star-formation rate, compared to normal galaxies. Whereas our Milky Way is forming stars with a rate of ∼ 3M ∕yr, the star-formation rate in starburst galaxies can be larger by a factor of more than a hundred. Dust heated by hot stars radiates in the FIR, rendering starbursts very strong FIR emitters. Many of them were discovered by the IRAS satellite (‘IRAS galaxies’) ; they are also called ULIRGs (Ultra Luminous InfraRed Galaxies).
The reason for this strongly enhanced star formation is presumably the interaction with other galaxies or the result of merger processes, an impressive example of which is the merging galaxy pair known as the ‘Antennae’ (see Fig. 9.25). In this system, stars and star clusters are currently being produced in very large numbers. The images show a large number of star clusters with a characteristic mass of 105 M , some of which are spatially resolved by the HST. Furthermore, particularly luminous individual stars (supergiants) are also identified. The ages of the stars and star clusters span a wide range and depend on the position within the galaxies. For instance, the age of the predominant population is about 5–10 Myr, with a tendency for the youngest stars to be located in the vicinity of strong dust absorption. However, stellar populations with an age of 100 and 500 Myr, respectively, have also been discovered; the latter presumably originates from the time of the first encounter of these two galaxies, which then led to the ejection of the tidal tails. This seems to be a common phenomenon; for example, in the starburst galaxy Arp 220 (see Fig. 1.​15) one also finds star clusters of a young population with age ≲ 107 yr, as well as older ones with age ∼ 3 × 108 yr. It thus seems that during the merging process several massive bursts of star-cluster formation are triggered.
/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig26_HTML.jpg
Fig. 9.26
The Antennae galaxies: In the left panel, superposed on the optical HST image are contours of infrared emission at 15 μm, measured by the ISO satellite. The right panel shows a composite: X-ray emission as observed by Chandra shown in blue, and the 8 μm map obtained by Spitzer (red), is superposed on an optical three-band composite image from HST. The strongest IR emission originates in optically dark regions. A large fraction of the star formation in this galaxy pair is not visible on optical images, because it is hidden by dust absorption. Note that the orientation in the right panel differs from that of the other images of the Antennae, as it was taken with the HST at a different orientation angle. Credit: Left: Canadian Astronomy and Astrophysics in the 21st Century, Courtesy Christine Wilson, McMaster University. Right: NASA/CXC/SAO/JPL-Caltech/STScI
It was shown by the ISO satellite that the most active regions of star formation are not visible on optical images, since they are completely enshrouded by dust (left panel in Fig. 9.26). A map at 8 μm obtained by the Spitzer Space Observatory (Fig. 9.26, right panel) shows the hot dust heated by young stars, where this IR emission is clearly anti-correlated with the optical radiation. Indeed, the mid-infrared emission is strongest in the region between these two colliding galaxies; apparently, this is the location where the current star formation is most intense. Maps of the CO emission also show that these regions contain a large reservoir of molecular gas. Furthermore, the large-scale X-ray radiation in this galaxy collision shows the efficiency with which gas is heated to high temperatures by the supernova events that accompany massive star formation, with a corresponding chemical enrichment of the gas with α-elements. Many of the X-ray point sources are due to high-mass X-ray binaries. Obviously, a complete picture of star formation in such galaxies can only be obtained from a combination of optical and IR images.
Combining deep optical and NIR photometry with MIR imaging from the Spitzer telescope, star-forming galaxies at high redshifts can be detected even if they contain an appreciable amount of dust (and thus may fail to satisfy the LBG selection criteria). These studies find that the comoving number density of ULIRGs with L IR ≳ 1012 L at z ∼ 2 is about three orders of magnitude larger than the local ULIRG density. These results seem to imply that the high-mass tail of the local galaxy population with M ≳ 1011 M was largely in place at redshift z ∼ 1. 5 and evolves passively from there on. We shall come back to this aspect below.
A129044_2_En_9_Fig27_HTML.gif
Fig. 9.27
Ultra-luminous compact X-ray sources (ULXs) in starburst galaxies. Upper left: The discrete X-ray sources in the Antenna galaxies; the size of the image is 4′ × 4′. Lower left: Optical (image) and (inlaid) Chandra image of the starburst galaxy NGC 253. Four of the ULXs are located within one kiloparsec from the center of the galaxy. The X-ray image is 2.​​2 × 2.​​2. Upper right: 5′ × 5′ Chandra image of the starburst galaxy M82; the diffuse radiation (red) is emitted by gas at T ∼ 106 K which is heated by the starburst and flows out from the central region of the galaxy. It is supposed that M82 had a collision with its companion M81 (see Fig. 6.​9) within the last 108 yr, by which the starburst was triggered. Lower right: The luminosity function of the ULXs in some starburst galaxies. The differently shaded regions indicate ranges in luminosity which correspond to Eddington luminosities of neutron stars, ‘normal’ stellar mass black holes, and black holes with M ≥ 10M . Credit: Upper left: NASA/SAO/CXC/G.Fabbiano et al. Bottom left: X-ray: NASA/SAO/CXC, optical: ESO. Upper right: NASA/SAO/G.Fabbiano et al., Bottom right: NASA/SAO/G.Fabbiano et al.
Ultra-luminous Compact X-ray Sources. Observations with the Chandra satellite have shown that starburst galaxies contain a rich population of very luminous compact X-ray sources (Ultra-luminous Compact X-ray Sources, or ULXs; see Fig. 9.27). They are formally defined to have an X-ray luminosity > 1039 erg∕s. Similar sources, though with lower luminosity, are also detected in the Milky Way, where these are binary systems with one component being a compact star (white dwarf, neutron star, or black hole). The X-ray emission is caused by accretion of matter (which we discussed in Sect. 5.​3.​2) from the companion star onto the compact component, and are called X-ray binaries.
Some of the ULXs in starbursts are so luminous, however, that the required mass of the compact star by far exceeds 1 M if the Eddington luminosity is assumed as an upper limit for the luminosity [see (5.​25)]. The detection of these ULXs in the 1980s by the Einstein observatory thus came unexpectedly, since one does not expect to form black holes with a mass larger than ∼ 10M in supernova explosions. Thus, one concludes that either the emission of these sources is highly anisotropic, hence beamed towards us, or that the sources are black holes with masses of up to ∼ 200M . In the latter case, we may just be witnessing the formation of intermediate mass black holes in these starbursts. In fact, recently a ULX was found with a peak luminosity of 1042 erg∕s, which, assuming that the Eddington limit is not exceeded, implies a black hole mass of > 500M as the origin of this source, located ∼ 4 kpc away from the center of an edge-on spiral galaxy.
This latter interpretation is also supported by the fact that the ULXs are concentrated towards the center of the galaxies—hence, these black holes may spiral into the galaxy’s center by dynamical friction, and there merge to a SMBH. This is one of the possible scenarios for the formation of SMBHs in the cores of galaxies, a subject to which we will return in Sect. 10.​4.​5. Furthermore, the similarity of the properties of ULXs with those of X-ray binaries may indicate that ULXs are just scaled-up version of these more common sources.

9.3.2 Extremely Red Objects (EROs)

As mentioned several times previously, the population of galaxies detected in a survey depends on the selection criteria. Thus, employing the Lyman-break method, one mainly discovers those galaxies at high redshift which feature active star formation and therefore have a blue spectral distribution at wavelengths longwards of Lyα. The development of NIR detectors enabled the search for galaxies at longer wavelengths. Of particular interest here are surveys of galaxies in the K-band, the longest wavelength window that is reasonably accessible from the ground (with the exception of the sub-millimeter to radio domain).
/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig28_HTML.jpg
Fig. 9.28
Redshift distribution of galaxies with K s < 20, as measured in the K20-survey. The shaded histogram represents galaxies for which the redshift was determined solely by photometric methods. The bin at z < 0 contains those nine galaxies for which it has not been possible to determine z. The peak at z ∼ 0. 7 is produced by two clusters of galaxies in the fields of the K20-survey. Source: A. Cimatti et al. 2002, The K20 survey. IV. The redshift distribution of K s < 20 galaxies: A test of galaxy formation models, A&A 391, L1, p. L2, Fig. 1. ©ESO. Reproduced with permission
/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig29_HTML.jpg
Fig. 9.29
Color-magnitude diagram, i.e., RK as a function of K, for sources in ten fields around clusters of galaxies. Most of the objects, in particular at the bright end, are located on the red sequence of early-type galaxies. We see that for faint magnitudes (roughly K ≥ 19), a population of sources with a very red color (about RK ≥ 5. 3) turns up. These objects are called EROs. Source: G.P. Smith 2002, A Hubble Space Telescope Survey of X-ray Luminous Galaxy Clusters: Gravitationally Lensed Arcs and EROs, astro-ph/0201236, Fig. 2. Reproduced by permission of the author
The NIR waveband is of particular interest because the luminosity of galaxies at these wavelengths is not dominated by young stars. As we have seen in Fig. 3.​34, the luminosity in the K-band depends rather weakly on the age of the stellar population, so that it provides a reliable measure of the total stellar mass of a galaxy.
Characteristics of EROs. Examining galaxies with a low K-band flux, one finds either galaxies with low stellar mass at low redshifts, or galaxies at high redshift with high optical (due to redshift) luminosity. But since the luminosity function of galaxies is relatively flat for L ≲ L , one expects the latter to dominate the surveys, due to the larger volume at higher z. In fact, K-band surveys detect galaxies with a broad redshift distribution. In Fig. 9.28 the z-distribution of galaxies in the K20-survey is shown. In this survey, objects with K s < 20 were selected in two fields with a combined area of 52 arcmin2, where Ks is a filter at a wavelength slightly shorter than the classic K-band filter. After excluding stars and Type 1 AGNs, 489 galaxies were found, 480 of which have had their redshifts determined. The median redshift in this survey is z ≈ 0. 8.
Considering galaxies in a (RK) vs. K color-magnitude diagram (Fig. 9.29), one can identify a population of particularly red galaxies, thus those with a large RK. These objects were named Extremely Red Objects (EROs); about 10 % of the galaxies in K-selected surveys at faint magnitudes are EROs, typically defined by RK > 5. Spectroscopic analysis of these galaxies poses a big challenge because an object with K = 20 and RK > 5 necessarily has R > 25, i.e., it is extremely faint in the optical domain of the spectrum. However, with the advent of 10-m class telescopes, spectroscopy of these objects has become possible.
The nature of EROs: passive ellipticals versus dusty starbursts. From these spectroscopic results, it was found that the class of EROs contains rather different kinds of sources. To understand this point we will first consider the possible explanations for a galaxy with such a red spectral distribution. As a first option, the object may be an old elliptical galaxy with the 4000 Å-break redshifted to the red side of the R-band filter, i.e., typically an elliptical galaxy at z ≳ 0. 8. For these galaxies to be sufficiently red to satisfy the selection criterion for EROs, they need to already contain an old stellar population by this redshift, which implies a very high redshift for the star formation episode in these objects; it is estimated from population synthesis models that their formation redshift must be z form ≳ 2. 5. A second possible explanation for large RK is reddening by dust. Such EROs may be galaxies with active star formation where the optical light is strongly attenuated by dust extinction. If these galaxies are located at a redshift of z ∼ 1, the measured R-band flux corresponds to a rest-frame emission in the UV region of the spectrum where extinction is very efficient.
Spectroscopic analysis reveals that both types of EROs are roughly equally abundant. Hence, about half of the EROs are elliptical galaxies that already have, at z ∼ 1, a luminosity similar to that of today’s ellipticals, and are at that epoch already dominated by an old stellar population. The other half are galaxies with active star formation which do not show a (prominent) 4000 Å-break but which feature the emission line of [Oii] at λ = 3727 Å, a clear sign of recent star formation. Further analysis of EROs by means of very deep radio observations confirms the large fraction of galaxies with high star-formation rates. Utilizing the close relation of radio emissivity and FIR luminosity, we find a considerable fraction of EROs to be ULIRGs at z ∼ 1.
Spatial correlations. EROs are very strongly correlated in space. The interpretation of this strong correlation may be different for the passive ellipticals and for those with active star formation. In the former case the correlation is compatible with a picture in which these EROs are contained in clusters of galaxies or in overdense regions that will collapse to a cluster in the future. The correlation of the EROs featuring active star formation can probably not be explained by cluster membership, but the origin of the correlation may be the same as for the correlation of the LBGs.
The number density of passive EROs, thus of old ellipticals, is surprisingly large compared with expectations from the model of hierarchical structure formation that we will discuss in Chap. 10.

9.3.3 Dusty star-forming galaxies

/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig30_HTML.jpg
Fig. 9.30
Spectral energy distribution of some dusty galaxies with known redshift z (symbols), together with two model spectra (curves). The spectral distributions for all sources are normalized to unity at λ = 850 μm. Four types of galaxies are distinguished: (I) IRAS galaxies at low z; (S) luminous sub-mm galaxies; (L) distant sources that are magnified by the gravitational lens effect and multiply imaged; (H) AGNs. Only a few sources among the lens systems (presumably due to differential magnification) and the AGNs deviate significantly from the model spectra. Source: A. Blain et al. 1999, Submillimeter-selected galaxies, astro-ph/9908111, Fig. 3. Reproduced by permission of the author
FIR emission from hot dust is one of the best indicators of star formation. However, observations in this waveband are only possible from space, such as was done with the IRAS and ISO satellites, and more recently with Spitzer and Herschel. Depending on the dust temperature, dust emission has its maximum at about 100 μm, which is not observable from the ground. At longer wavelengths there are spectral windows longwards of λ ∼ 250 μm where observations through the Earth’s atmosphere are possible, for instance at 450 and 850 μm in the sub-millimeter waveband. However, the observing conditions at these wavelengths are extremely dependent on the amount of water vapor in the atmosphere, so that the observing sites must by dry and at high elevations. In the sub-millimeter (sub-mm) range, the long wavelength domain of thermal dust radiation can be observed, which is illustrated in Fig. 9.30.
/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig31_HTML.jpg
Fig. 9.31
A sub-millimeter image of the COSMOS field, taken with the SPIRE instrument on Herschel. The image is about 2 degrees on the side and is a color composite of three bands with wavelength 250, 350 and 500 μm. For six galaxies in the field, the zooms show an optical images, with the cosmic look-back time indicated as obtained by spectroscopy with the W.M. Keck telescopes. Credit: COSMOS field: ESA/Herschel/SPIRE/HerMES Key Programme; Hubble images: NASA, ESA; Keck Spectra: Caltech/W. M. Keck Observatory
Developments. Since about 1998 sub-mm astronomy has experienced an enormous boom, with two instruments having been put into operation: the Sub-millimeter Common User Bolometer Array (SCUBA), operating at 450 and 850 μm, with a field-of-view of 5 arcmin2, and the Max-Planck Millimeter Bolometer (MAMBO), operating at 1200 μm. Both are bolometer arrays which initially had 37 bolometers each, but which later were upgraded to a considerably larger number of bolometers. With the opening of the 12-m APEX (Atacama Pathfinder Experiment) telescope (see Fig. 1.​28) in Chile, equipped with powerful instrumentation, a further big step in sub-mm astronomy was achieved. In addition, the Herschel satellite (Fig. 1.​33), operating at wavelength between 55 and 670 μm, allowed imaging of large sky areas at these wavelengths, due to a much lower noise level than can be achieved from the ground, though with considerably worse spatial resolution due to the smaller aperture compared to ground-based sub-mm telescopes. Figure 9.31 shows a Herschel image of the COSMOS field.
Apart from single-dish observatories, interferometers operating at these wavelengths yield substantially higher angular resolution, for example the very successful IRAM Plateau de Bure interferometer in the French Alps consisting of six 15-m antennas. The Atacama Large Millimeter Array (ALMA; see Fig. 1.​29) with its 54 12-m and 12 7-m antennas, inaugurated in 2013, marks a huge leap in terms of resolution and sensitivity in this waveband regime.
/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig32_HTML.jpg
Fig. 9.32
Predicted flux from dusty galaxies as a function of redshift. The bolometric luminosity of these galaxies is kept constant. The solid red and the blue dashed curves show the flux at λ = 850 μm and λ = 175 μm, respectively. On the right, the index β of the dust emissivity is varied, and the temperature of the dust T d = 38 K is kept fixed. On the left, β = 1. 5 is fixed and the temperature is varied. It is remarkable how flat these curves are over a very wide range in redshift, in particular at 850 μm; this is due to the very strong negative K-correction which derives from the spectral behavior of thermal dust emission, shown in Fig. 9.30. Whereas for these model calculations an Einstein–de Sitter model was assumed, the behavior is very similar also in Λ-dominated universes. Source: A. Blain et al. 1999, Submillimeter-selected galaxies, astro-ph/9908111, Fig. 1. Reproduced by permission of the author
The negative K-correction of sub-mm sources. The emission of dust at these wavelengths is described by a Rayleigh–Jeans spectrum, modified by an emissivity function that depends on the dust properties (chemical composition, distribution of dust grain sizes); typically, one finds
 $$\displaystyle{S_{\nu } \propto \nu ^{2+\beta }\quad \mathrm{with}\quad \beta \sim 1\ldots 2\;.}$$
This steep spectrum for frequencies below the peak of the thermal dust emission at λ ∼ 100 μm implies a very strong negative K-correction (see Sect. 5.​6.​1) for wavelengths in the sub-mm domain: at a fixed observed wavelength, the rest-frame wavelength becomes increasingly smaller for sources at higher redshift, and there the emissivity is larger. As Fig. 9.32 demonstrates, this spectral behavior causes the effect that the flux in the sub-mm range does not necessarily decrease with redshift. For z ≲ 1, the 1∕D 2-dependence of the flux dominates, so that up to z ∼ 1 sources at fixed luminosity get fainter with increasing z. However, between z ∼ 1 and z ∼ z flat the sub-mm flux as a function of redshift remains nearly constant or even increases with z, where z flat depends on the dust temperature T d and the observed wavelength; for T d ∼ 40 K and λ ∼ 850 μm one finds z flat ∼ 8. We therefore have the quite amazing situation that sources appear brighter when they are at larger distances. This is caused by the very negative K-correction which more than compensates for the 1∕D 2-decrease of the flux. Only for z > z flat does the flux begin to rapidly decrease with redshift, since then, due to redshift, the corresponding restframe frequency is shifted to the far side of the maximum of the dust spectrum (see Fig. 9.30). Hence, a sample of galaxies that is flux-limited in the sub-mm domain should have a very broad z-distribution. The dust temperature is about T d ∼ 20 K for low-redshift spirals, and T d ∼ 40 K is a typical value for galaxies at higher redshift featuring active star formation. The higher T d, the smaller the sub-mm flux at fixed bolometric luminosity.
/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig33_HTML.jpg
Fig. 9.33
The sub-mm galaxy SMM J09429+4658. The three images on the right have a side length of 30″ each, centered on the center of the error box of the 850 μm observation. The smaller image on the left is the difference of two HST images in red and infrared filters, showing the dust disk in the spiral galaxy H1. The second image from the left displays an R-band image, superposed with the contours of the SCUBA 850 μm emission. The second image from the right is an I-band image, superposed with the contours of radio emission at 1. 4 GHz, and the right-most panel shows a K-band image. The radio contours show emission from the galaxy H1 (z = 0. 33), but also weaker emission right at the center of the sub-mm map. In the K-band, a NIR source (H5) is found exactly at this position. It remains unclear which of these two sources is the sub-mm source, but the ratio of sub-mm to 1. 4 GHz emission would be atypical if H1 is identified with the sub-mm source. Source: A. Blain et al. 1999, Submillimeter-selected galaxies, astro-ph/9908111, Fig. 4. Reproduced by permission of the author
Counts of sub-mm sources at high Galactic latitudes have yielded a far higher number density than was predicted by early galaxy evolution models. For the density of sources as a function of limiting flux S, at wavelength λ = 850 μm, one obtains
 $$\displaystyle{ N(> S) \simeq 7.9 \times 10^{3}\left ( \frac{S} {1\,\mathrm{mJy}}\right )^{-1.1}\,\mathrm{deg}^{-2}\;. }$$
(9.1)
The identification of sub-mm sources. At first, the optical identification of these sources turned out to be extremely difficult: due to the relatively low angular resolution of single-dish sub-millimeter telescopes (for example, MAMBO has a beam with FWHM of ∼ 11″ at λ = 1. 2 mm), the positions of sources can only be determined with an accuracy of several acrseconds.5 Typically, several faint galaxies can be identified on deep optical images within an error circle of this radius. Furthermore, Fig. 9.32 suggests that these sources have a relatively high redshift, thus they should be very faint in the optical. An additional problem is reddening and extinction by the same dust that is the source of the sub-mm emission.
The identification of sub-mm sources was finally accomplished by means of their radio emission, since a majority of the sources selected at sub-mm wavelengths can be identified in very deep radio observations at 1. 4 GHz. Since the radio sky is far less crowded than the optical one, and since the VLA achieves an angular resolution of ∼ 1″ at λ = 20 cm, the optical identification of the corresponding radio source becomes relatively easy. One example of this identification process is shown in Fig. 9.33. With the accurate radio position of a sub-mm source, the optical identification can then be performed. In most cases, they are very faint optical sources indeed, so that spectroscopic analysis is difficult and very time-consuming. Another method for estimating the redshift results from the spectral energy distribution shown in Fig. 9.30. Since this spectrum seems to be nearly universal, i.e., not varying much among different sources, some kind of photometric redshift can be estimated from the ratio of the fluxes at 1.4 GHz and 850 μm, yielding reasonable estimates in many cases.
Redshift distribution of SMGs. The median redshift of sources with 850 μm flux larger than 5 mJy and 1. 4 GHz flux above 30 μJy is about 2.2. However, at these sensitivities about half the sub-mm sources are unidentified in the radio, and hence their redshift distribution could be different from those with radio counterparts. In some of the sources with radio identification, an AGN component, which contributes to the dust heating, was identified, but in general newly born stars seem to be the prime source of the energetic photons which heat the dust. The optical morphology and the number density of the sub-mm sources suggest that we are witnessing the formation of massive elliptical galaxies in these sub-mm sources.
The great capabilities of interferometric observations enables a different method for identification and redshift determination of SMGs. The high angular resolution of the Plateau de Bure and, in particular, ALMA interferometers can pinpoint a sub-mm source very accurately, allowing the identification with optical or infrared sources. In addition, the large bandwidth of the ALMA receivers can take spectra of these sources over an appreciable range of wavelengths, and thus identify emission lines of molecules, predominantly those of the CO and water molecules, or fine-structure lines of atoms (specifically, ionized and neutral carbon), therefore removing the need for obtaining optical spectra of these optically faint sources. The potential of this method was recently established: even by observing with an incomplete array of telescopes, ALMA detected emission lines in 23 out of 26 sources selected by the South Pole Telescope (SPT; see Fig. 1.​31) at 1. 4 mm wavelength.
Once the precise locations of the sub-mm sources are determined with interferometric observations, they can be identified on deep optical images, and their redshifts be determined from optical spectroscopy. Hence, much of the potential bias in the redshift distribution of these sources, which are caused by the fact that only about half of the SMGs have a radio identification, are removed.6 Indeed, if one compares the redshift distributions of both samples, shown in Fig. 9.34, one sees that they are significantly different. Using radio-identification as an intermediate step, and neglecting those SMGs for which no radio source could be identified, biases the redshift distribution of dusty star-forming galaxies low. The population extends over a significantly larger redshift range than concluded previously—in accord with expectations from the large negative K-correction. In addition, the redshifts of SMGs can be determined with the PdB and ALMA interferometers directly, using molecular line spectroscopy, without the need for optical spectroscopy. In particular the large bandwidth of the ALMA receiver allows one to cover a broad range of wavelengths, and thus of redshifts for which molecular lines can be detected and identified.
A129044_2_En_9_Fig34_HTML.gif
Fig. 9.34
The grey line shows the cumulative redshift distribution of dusty star-forming galaxies selected from the SPT survey, as determined by molecular line spectroscopy with ALMA; the distribution has a median redshift of z med ∼ 3. 5. The blue curve is the redshift distribution of SMGs that were identified with radio sources and subsequently their redshift was determined by optical spectroscopy, yielding z med ∼ 2. 2. The orange curve is the redshift distribution of SMGs identified at mm-wavelengths, with redshifts determined mainly through photometric fitting. Source: J.D. Vieira et al. 2013, Dusty starburst galaxies in the early Universe as revealed by gravitational lensing, Nature 495, 344, Fig. 3. Reprinted by permission of Macmillan Publishers Ltd: Nature, ©2013
Halo masses of sub-mm galaxies. As we discussed in Sect. 8.​1.​2, one can estimate the mass of the dark matter halo in which objects reside, by comparing their clustering properties with those of dark matter halos, as obtained from cosmological simulations. Sub-mm galaxies identified in a survey conducted by APEX (see the left panel of Fig. 9.35) allowed an estimate of the clustering length r 0 (defined such that the two-point correlation function is unity at separation r 0) of these sources, yielding r 0 ≈ (7. 5 ± 2)h −1 Mpc. In the right panel of Fig. 9.35, this measurement of r 0 is related to that of other source populations. The clustering length of sub-mm sources is very similar to that of QSOs at the same redshift, and considerably larger than that for Lyman-break galaxies. Comparing this to the clustering length of dark matter halos with different masses, shown as dotted curves in Fig. 9.35, we conclude that SMGs live in relatively massive halos of several times 1012 M at z ∼ 2. This results was recently confirmed by detecting a weak lensing signal around a sample of ∼ 600 relatively bright SMGs, which yields a characteristic halo mass of these galaxies of ∼ 1013 M . This mass is comparable to that of current epoch massive elliptical galaxies, and suggests the interpretation that high-redshift SMGs evolve into present-day ellipticals. A large fraction of their stellar population is formed in the epoch at which the galaxy is seen as a SMG; by the end of this period, most of the gas in the galaxy is used up (or some fraction of it may be expelled), and in the remaining evolution little or no star formation occurs.
Additional support for this idea is provided by the fact that the sub-mm galaxies are typically brighter and redder than (restframe) UV-selected galaxies at redshifts z ∼ 2. 5. This indicates that the stellar masses in sub-mm galaxies are higher than those of LBGs.
A joint investigation of z ∼ 2 sub-mm galaxies at X-ray, optical and MIR wavelengths yields that these sources are not only forming stars at a high rate, but that they already contain a substantial stellar population with M ∼ 1011 M , roughly an order of magnitude more massive than LBGs at similar redshifts. The large AGN fraction among sub-mm galaxies indicates that the growth of the stellar population is accompanied by accretion and thus the growth of supermassive black holes in these objects. Nevertheless, the relatively faint X-ray emission from these galaxies suggests that either their SMBHs have a mass well below the local relation between M and stellar properties of (spheroidal) galaxies, or that they accrete at well below the Eddington rate. Furthermore, the typical ratio of X-ray to sub-mm luminosity of these sources is about one order of magnitude smaller than in typical AGNs, which seems to imply that the total luminosity of these sources is dominated by the star-formation activity, rather than by accretion power. This conclusion is supported by the fact that the optical counterparts of SMGs show strong signs of merging and interactions, together with their larger size compared to optically-selected galaxies at the same redshifts. This latter point shows that the emission comes from an extended region, as expected from star formation in mergers, rather than AGN activity.
A129044_2_En_9_Fig35_HTML.gif
Fig. 9.35
Left panel: Sub-millimeter sources in the Extended Chandra Deep Field South (ECDFS). The image shows highly significant detections of sub-mm sources at 870 μm, as observed with the 12-m APEX telescope, superimposed on a Spitzer mid-IR image of the same sky region. On the 0. 35 deg2 of the ECDFS, 126 sub-mm sources were detected with a significance higher than 3. 7σ. Right panel: Based on a clustering analysis of the sub-mm sources shown on the left, the clustering length r 0 of the sources was determined. The figure shows the clustering length as a function of redshift, for different types of sources: QSOs over a broad range of redshifts, local ellipticals and luminous red galaxies, local blue galaxies and star-forming galaxies at intermediate redshift, as selected by Spitzer at 24 μm wavelength, high-redshift Lyman-break galaxies, and galaxy clusters. The dotted curves indicate the clustering length of dark matter halos as a function of redshift, for different halo masses as indicated. Credit: Left: ESO, APEX (MPIfR/ESO/OSO), A. Weiss et al., NASA Spitzer Science Center. Right: R.C. Hickox et al. 2012, The LABOCA survey of the Extended Chandra Deep Field-South: clustering of submillimetre galaxies, MNRAS 421, 284, p. 291, Fig. 6b. Reproduced by permission of Oxford University Press on behalf of the Royal Astronomical Society
/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig36_HTML.jpg
Fig. 9.36
The sub-millimeter galaxy HXMM01, detected by Herschel in the framework of the HerMES survey. In blue, a K-band image shows the presence of four main sources in this field; the two big ones are galaxies at z = 0. 66 (left) and z = 0. 50 (right). The two smaller ones, separated by ∼ 3″, are at the same redshift of z = 2. 31. In red, a sub-millimeter maps is shown, whereas green displays molecular emission. The two yellow spots show the regions where near-IR, sub-mm and molecular emission are spatially coincident; from those regions, the bulk of the flux seen by Herschel is emitted. The image has a size of about 10″ on a side. Credit: ESA/NASA/JPL-Caltech/UC Irvine/STScI/Keck/NRAO/SAO
Merging SMGs. The Herschel satellite found a strong sub-mm source, whose subsequent reobservations with different telescopes revealed this to be a merger of two SMGs. This source, called HXMM01, is shown in Fig. 9.36. In the near-IR, the image shows four main sources, of which the two brightest ones are galaxies at intermediate redshifts. The two fainter ones are galaxies at z = 2. 31. Those coincide with strong sub-mm sources, as seen by the interferometric images at 880 μm, as well as through their strong molecular line emission. This pair of sources has a separation of 19 kpc, and is most likely undergoing a merger. The two foreground galaxies will cause a moderate lensing magnification of the SMGs, with an estimated μ ∼ 1. 6. Thus, even after magnification correction, this source belongs to the brightest SMGs, with a corrected flux of ∼ 20 mJy at λ = 850 μm. This source is extreme in several properties: Its dust temperature is estimated to be T ≈ 55 K, which is larger than that of most starburst galaxies. The estimated bolometric infrared luminosity is L ∼ 2 × 1013 L , with a corresponding star-formation rate of  $$\dot{M} \sim 2000M_{\odot }/\mathrm{yr}$$ . At this rate, the molecular gas of ∼ 2. 3 × 1011 M will be turned into stars on a timescale of only 70 Myr. The brevity of this time interval implies that objects similar to HXMM01 should be rare. The molecular gas mass in this object is comparable with the stellar mass already present, i.e., the gas-mass fraction in this object is about 50 %. At the end of its merging process and star-formation episode, the resulting galaxy will have a stellar mass of ∼ 1011 M , i.e., the stellar mass of a massive elliptical galaxy.
Although HXMM01 is not the only pair of merging SMGs yet discovered, it is brighter than the other ones by a factor of ∼ 2 (magnification corrected); correspondingly, its star-formation rate is also about twice that of the other merging systems. There are indications that the far-IR luminosity, and thus the star-formation rate, at fixed molecular mass (as measured by the luminosity in molecular CO lines), is higher by about a factor of three compared to galaxies that are not observed to be merging. This then is a direct hint at the possibility that merging triggers star-formation events, or bursts of star-formation.
In fact, there are several observations which suggest that the very large star-forming rates of many of the most luminous SMGs are due to merging events. Arguably the most direct one comes from interferometric observations of the molecular gas in SMGs, which shows that most of them have morphological and kinematical properties that identify them as merging systems. Hence, in accordance with the high merger rate of local ULIRGs, the extreme SMGs may also be triggered by merger events.
If dusty galaxies are at very high redshifts z ≳ 5, the peak of their spectral energy distribution shown in Fig. 9.30 is observed at wavelengths longward of 500 μm. Hence, for such sources the flux density is expected to increase with wavelength for λ ≲ 500 μm. The SPIRE instrument on the Herschel Space Observatory observed at 250, 350 and 500 μm, and blank survey fields at these frequencies can thus be used to search for very high-redshift sub-millimeter galaxies.
One such object found by this selection method is shown in Fig. 9.37. The galaxy HFLS3 is an extreme starburst at z = 6. 34, with an estimated star-formation rate of ∼ 3000 M ∕yr, some factor of 20 larger than the local starburst Arp 220 (see Fig. 1.​15). With an estimated gas mass of ∼ 1011M , this object would transform all its gas mass into stars on a time-scale of only ∼ 30 Myr, assuming a constant star-formation rate. Spatially resolved molecular spectroscopy shows a velocity profile of this galaxy, which can be used to estimate a dynamical mass of ∼ 2. 7 × 1011 M , yielding a gas-mass fraction of ∼ 40 %, similar to what is found in typical z ∼ 2 sub-millimeter galaxies.
Magnification bias of sub-millimeter sources. We mentioned before that some of the apparently most luminous sources of any kind have a large probability of being gravitationally lensed. The reason for this can be seen as follows: If we consider a population of sources, then their distribution in luminosity is most frequently described by a Schechter-like function. That means that for low luminosities, the luminosity function behaves as a power law in L, whereas for L > L , where L characterizes the break in the luminosity function, the density of sources decreases exponentially with L. In particular that means that there are essentially no sources with luminosity L ≳ 5L .
The probability that any given high-redshift source is gravitationally lensed and significantly magnified is small, of order 10−4. Thus, if one picks random sources, the fraction of lensed ones among them will be similarly small. However, we can not pick random sources, but only sources above the flux limit of our observations. The situation in flux-limited samples can be quite different, since the magnification by lensing affects the mix of sources which are above the flux threshold. The probability that a source undergoes a magnification larger than μ can be shown to behave like μ −2, provided the source is sufficiently small. Hence, large magnifications are correspondingly rare. But the probability for large magnifications decreases as a power law in μ—compared to the exponential decrease in the luminosity of sources for L > L . There will be a point where the (low-amplitude) power law overtakes the exponential, or in other words, where a source above a given flux threshold is more likely to be highly magnified, than having a very high intrinsic luminosity. An example of this effect are the bright z ∼ 3 Lyman-break galaxies shown in Figs. 9.17 and 9.18, whose inferred luminosity is much larger than the L of this population of galaxies.
/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig37_HTML.jpg
Fig. 9.37
The galaxy HFLS3 at z = 6. 34, found by a color selection in the sub-millimeter regime. The big image shows a part of the HerMES blank field survey, with cut-outs at different wavelengths shown as small panels on the right, as well as an optical image and one combining near-IR and millimeter imaging. Credit: ESA/Herschel/HerMES/ IRAM/GTC/W.M. Keck Observatory
This is illustrated in Fig. 9.38, where the upper end of the 500 μm source counts are shown, together with a model decomposition of the counts into lensed and unlensed SMGs, nearby spiral galaxies and radio AGNs. As we just argued, for relatively low fluxes, the fraction of lensed sources is increasingly small. However, due to the steep decline of the unlensed counts, beyond a certain flux level they start to dominate the counts of SMGs. From that figure, one therefore expects that a SMG with S ≳ 60 mJy at 500 μm has a fairly high probability of being lensed, whereas a SMG with S ≳ 100 mJy almost certainly is a lensed source.
Hence, selecting sources with S ≳ 100 mJy, and cleaning the source catalog for nearby spirals (they are bright and big in the optical, and can thus easily be identified), one expects to obtain a clean lensed sample. Indeed, all five candidates selected from the survey data on which Fig. 9.38 is based, are most likely lensed, with the putative lens being identified on optical and NIR images. Furthermore, interferometric observations have confirmed the lensing nature of several of the candidates, by finding multiple images.
/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig38_HTML.jpg
Fig. 9.38
The dominance of gravitationally-lensed magnified sub-mm galaxies at the bright end. The dots show the upper end of the 500 μm source counts of SMGs, obtained from the Herschel ATLAS survey. The red curve shows a model of the unlensed source counts, which has the shape of a Schechter-function, and drops off exponentially for large S. The red dashed curve shows the corresponding counts of lensed and magnified sources, whereas the blue curve is the contribution from low-redshift spiral galaxies. The lensed SMGs overtake the unlensed ones for fluxes ≳ 60 mJy. Source: M. Negrello et al. 2010, The Detection of a Population of Submillimeter-Bright, Strongly-Lensed Galaxies, arXiv:1011.1255, Fig. 1. Reproduced by permission of the author
Using ALMA imaging of sources from the South Pole Telescope survey, selected at 1.4 and 2 mm to have the spectral index at these wavelengths corresponding to dust, and cleaning the sample for nearby objects (e.g., excluding sources detected by IRAS or in low-frequency radio catalogs), a large fraction of these sources turn out to be gravitationally lensed. In Fig. 9.39, ten of these are shown, where not only the multiple images of the sources (which are all at high redshift) are shown, but also a near-IR image of the field which clearly shows the lensing galaxy. Hence, selecting bright sources in the (sub)-millimeter regime yields a very high success rate of finding gravitational lens systems. All these lens systems would be missed in optical surveys, due to the faintness of the SMGs at optical and NIR wavelengths.
/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig39_HTML.jpg
Fig. 9.39
Near-infrared (greyscale) and ALMA 870 μm images of ten high flux millimeter-selected sources from the South Pole Telescope survey. The redshifts of the SMGs are indicated in the lower left corner of each panel. All ten sources are obviously gravitationally lensed by a foreground galaxy. Source: J.D. Vieira et al. 2013, Dusty starburst galaxies in the early Universe as revealed by gravitational lensing, arXiv:1303.2723, Fig. 1. Reproduced by permission of the author

9.3.4 Damped Lyman-alpha systems

In our discussion of QSO absorption lines in Sect. 5.​7, we mentioned that the Lyα lines are broadly classed into three categories: the Lyα forest, Lyman-limit systems, and damped Lyα systems, which are separated by a column density of N HI ∼ 1017 cm−2 and N HI ∼ 2 × 1020 cm−2, respectively. The origin of the Lyα forest, as discussed in some detail in Sect. 8.​5, is diffuse, highly ionized gas with small density contrast. In comparison, the large column density of damped Lyα systems (DLAs) strongly suggests that hydrogen is mostly neutral in these systems. The reason for this is self-shielding : for column densities of N HI ≳ 2 × 1020 cm−2 the background of ionizing photons is unable to penetrate deeply into the corresponding hydrogen ‘cloud’, so that only its surface is highly ionized. Interestingly enough, this column density is about the same as that observed in 21 cm hydrogen emission at the optical radius of nearby spiral galaxies.
DLAs can be observed at all redshifts z ≲ 5. For z > 5 the Lyα forest becomes so dense that these damped absorption lines are very difficult to identify. For z ≲ 1. 6 the Lyα transition cannot be observed from the ground; since the apertures of optical/UV telescopes in space are considerably smaller than those on the ground, low-redshift DLAs are substantially more complicated to observe than those at higher z.
The neutral hydrogen mass contained in DLAs. The column density distribution of Lyα forest lines is a power law, given by (8.​30). The relatively flat slope of β ∼ 1. 6 indicates that most of the neutral hydrogen is contained in systems of high column density. This can be seen as follows: the total column density of neutral hydrogen above some minimum column density N min is
 $$\displaystyle\begin{array}{rcl} N_{\mathrm{HI,tot}}(N_{\mathrm{max}})& \propto & \int _{N_{\mathrm{min}}}^{N_{\mathrm{max}} }\mathrm{d}N_{\mathrm{HI}}\;N_{\mathrm{HI}}\, \frac{\mathrm{d}N} {\mathrm{d}N_{\mathrm{HI}}} \\ & \propto & \int _{N_{\mathrm{min}}}^{N_{\mathrm{max}} }\mathrm{d}N_{\mathrm{HI}}\;N_{\mathrm{HI}}^{1-\beta } = \frac{N_{\mathrm{max}}^{2-\beta }- N_{\mathrm{ min}}^{2-\beta }} {2-\beta } \;,{}\end{array}$$
(9.2)
and is, for β < 2, dominated by the highest column density systems. In fact, unless the distribution of column densities steepens for very high N HI, the integral diverges. From the extended statistics now available for DLAs, it is known that dN∕dN HI attains a break at column densities above N HI ≳ 1021 cm−2, rendering the above integral finite. Nevertheless, this consideration implies that most of the neutral hydrogen in the Universe visible in QSO absorption lines is contained in DLAs. From the observed distribution of DLAs as a function of column density and redshift, the density parameter Ω HI in neutral hydrogen as a function of redshift can be inferred. Apparently, Ω HI ∼ 10−3 over the whole redshift interval 0 < z < 5, with perhaps a small redshift dependence. Compared to the current density of stars, this neutral hydrogen density is smaller only by a factor ∼ 3. Therefore, the hydrogen contained in DLAs is an important reservoir for star formation, and DLAs may represent condensations of gas that turn into ‘normal’ galaxies once star-formation sets in. Since DLAs have low metallicities, typically 1/10 of the Solar abundance, it is quite plausible that they have not yet experienced much star formation.
The nature of DLAs. This interpretation is supported by the kinematical properties of DLAs. Whereas the fact that the Lyα line is damped implies that its observed shape is essentially independent of the Doppler velocity of the gas, velocity information can nevertheless be obtained from metal lines. Every DLA is associated with metal absorption line systems, covering low- and high-ionization species (such as Siii and Civ, respectively) which can be observed by choosing the appropriate wavelength coverage of the spectrum. The profiles of these metal lines are usually split up into several components. Interpreted as ionized ‘clouds’, the velocity range Δ v thus obtained can be used as an indicator of the characteristic velocities of the DLA. The values of Δ v cover a wide range, with a median of ∼ 90 km∕s for the low-ionization lines and ∼ 190 km∕s for the high-ionization transitions. The observed distribution is largely compatible with the interpretation that DLAs are rotating disks with a characteristic rotational velocity of v c ∼ 200 km∕s, once random orientations and impact parameters of the line-of-sight to the QSO are taken into account.
Search for emission from DLAs. If this interpretation is correct, then we might expect that the DLAs can also be observed as galaxies in emission. This, however, is exceedingly difficult for the high-redshift DLAs. Noting that they are discovered as absorption lines in the spectrum of QSOs, we face the difficulty of imaging a high-redshift galaxy very close to the line-of-sight to a bright QSO (to quote characteristic numbers, the typical QSO used for absorption-line spectroscopy has B ∼ 18, whereas an L -galaxy at z ∼ 3 has B ∼ 24. 5). Due to the size of the point-spread function this is nearly hopeless from the ground. But even with the resolution of HST, it is a difficult undertaking. Another possibility is to look for the Lyα emission line at the absorption redshift, located right in the wavelength range where the DLA fully blocks the QSO light. However, as we discussed for LBGs above, not all galaxies show Lyα in emission, and it is not too surprising that these searches have largely failed. Only very few DLA have been detected in emission, with some of them seen only through the Lyα emission line at the trough of the damped absorption line, but with no observable continuum radiation. This latter fact indicates that the blue light from DLAs is considerably fainter than that from a typical LBG at z ∼ 3, consistent with the interpretation that DLAs are not strong star-forming objects. But at least one DLA is observed to be considerably brighter and seems to share some characteristics of LBGs, including a high star-formation rate. In addition, a couple of DLAs have been detected by [Oiii] emission lines. Overall, then, the nature of high-redshift DLAs is still unclear, due to the small number of direct identifications.
For DLAs at low redshifts the observational situation is different, in that a fair fraction of them have counterparts seen in emission. Whereas the interpretation of the data is still not unambiguous, it seems that the low-redshift population of DLAs may be composed of normal galaxies.
The spatial abundance of DLAs is largely unknown. The observed frequency of DLAs in QSO spectra is the product of the spatial abundance and the absorption cross section of the absorbers. This product can be compared with the corresponding quantity of local galaxies: the detailed mapping of nearby galaxies in the 21 cm line shows that their abundance and gaseous cross section are compatible with the frequency of DLAs for z ≲ 1. 5, and falls short by a factor ∼ 2 for the higher-redshifts DLAs.
/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig40_HTML.jpg
Fig. 9.40
Composite image of a Lyα blob at z = 3. 09. The yellow color shows the Lyα emission, obtained by a narrow-band filter exposure. Inside the blob a galaxy is located, as seen in an optical (white) and infrared (8 μm, red) broadband image. X-ray emission (shown in blue) indicates that the Lyα emission is powered by an AGN. This image has a size of 38″. Credit: X-ray (NASA/CXC/Durham Univ./D.Alexander et al.); Optical (NASA/ESA/STScI/IoA/S.Chapman et al.); Lyman-alpha Optical (NAOJ/Subaru/Tohoku Univ./T.Hayashino et al.); Infrared (NASA/JPL-Caltech/Durham Univ./J.Geach et al.)

9.3.5 Lyman-alpha blobs

The search for high-redshift galaxies with narrow-band imaging, where the filter is centered on the redshifted Lyα emission line, has revealed a class of objects which are termed ‘Lyman-α blobs’. These are luminous and very extended sources of Lyα emission; their characteristic flux in the Lyα line is ∼ 1044 erg∕s, and their typical size is ∼ 30 to ∼ 100 kpc. Some of these sources show no detectable continuum emission in any broad-band optical filter. Hence, these sources seem to form a distinct class from the Lyman-alpha emitters discussed previously.
The nature of these high-redshifts objects remained unknown for a long time. Suggested explanations were wide-ranging, including a hidden QSO, strong star formation and associated superwinds, as well as ‘cold accretion’, where gas is accreted onto a dark matter halo and hydrogen is collisionally excited in the gas of temperature ∼ 104 K, yielding the observed Lyα emission. It even seems plausible that the Lyman-α blobs encompass a range of different phenomena, and that all three modes of powering the line emission indeed occur. As a common feature, most of the Lyman-α blobs are associated with luminous galaxies, and are associated with strong infrared emission.
Chandra observations of 29 Lyα blobs detected five on them in X-rays; one of them is shown in Fig. 9.40. The X-ray sources are AGNs with L X ∼ 1044 erg∕s and rather large obscuration. Furthermore, these sources emit infrared light from warm dust. The energy output of the AGN is sufficiently large to power the Lyα emission through photoionization. Hence, the AGN hypothesis has been verified for at least some of the sources.
Two of these Lyα blobs were discovered by narrow-band imaging of the aforementioned proto-cluster of LBGs at z = 3. 09. Both of them are sub-mm sources and therefore star-forming objects; the more powerful one has a sub-mm flux suggesting a star-formation rate of ∼ 1000M ∕yr. Spatially resolved spectroscopy extending over the full ∼ 100 kpc size of one of the Lyα blobs shows that across the whole region there is an absorption line centered on the Lyα emission line. The optical depth of the absorption line suggests an Hi column density of ∼ 1019 cm−2, and its centroid is blueshifted relative to the underlying emission line by ∼ 250 km∕s. The spatial extent of the blueshifted absorption shows that the outflowing material is a global phenomenon in this object—a true superwind, most likely driven by energetic star formation and subsequent supernova explosions in these objects. Therefore, it appears likely that Lyα blobs are intimately connected to massive star-formation activity.

9.4 Properties of galaxies at high redshift

9.4.1 Demography of high-redshift galaxies

Being able to detect galaxies at high redshifts, we may first consider their abundance and investigate how their luminosity distribution compares with that galaxies in the local Universe. We are interested in a possible evolution of the luminosity function of galaxies with redshift, as this would clearly indicate that the galaxy population as a whole changes with redshift. The source counts from the Hubble Deep Field (Fig. 9.11) and its strong deviations from the non-evolution models provide a clear indication that the galaxy population evolves in redshift. This point will be considered here in somewhat more detail.
The most convenient way of summarizing the results is a representation of the estimated luminosity function in terms of a fit with a Schechter function (3.​52). In that, the luminosity function is characterized by an overall normalization Φ , a characteristic luminosity L (or, equivalently, an absolute magnitude M ) above which the abundance decreases exponentially, and a power-law slope α of the luminosity function at L ≪ L . An evolution of Φ with redshift indicates that the abundance of luminous galaxies evolves. If L depends on z, one may conclude that the luminosity of a typical galaxy is different at higher redshifts. Finally, the faint-end slope α determines what fraction of the total luminosity of the galaxy population is emitted by the fainter galaxies—see (3.​58).
The UV-luminosity function. Since high-redshift galaxies are selected using quite a variety of methods, as discussed in Sect. 9.1, and since the nature of the detected galaxies depends on their selection method, one has to consider different types of luminosity functions. For example, the Lyman-break method selects galaxies by their rest-frame UV radiation, so that from these surveys, the UV-luminosity function of galaxies can be obtained. Since the Lyman-break technique is applicable over a very wide redshift range, the UV-luminosity function has been obtained for redshifts between 2 and 7.
As already indicated by the galaxy counts shown in Fig. 9.11, the rest-frame UV-luminosity function of galaxies evolves strongly with redshift. In the redshift interval 2 ≲ z ≲ 4, the characteristic luminosity L is about three magnitudes brighter than that of the local UV-luminosity function as determined with GALEX. This immediately shows that a typical galaxy at these redshifts is far more actively forming stars than local galaxies. Furthermore, the faint-end slope α is steeper for the high-redshift galaxies than for local ones. In fact, estimates yield α ∼ −1. 6, indicating that much of the UV-luminosity density at high redshifts is emitted from rather faint galaxies—galaxies which are currently not observed due to the limited sensitivity of our instruments. Therefore, the overall abundance of UV-luminous galaxies is considerably larger in the redshift interval 2 ≲ z ≲ 4 than it is today. Since the UV-radiation is produced by massive (and thus young) stars, this implies that the star-formation activity at those redshifts was much more intense than at the current epoch.
/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig41_HTML.jpg
Fig. 9.41
High-redshift rest-frame UV luminosity functions of galaxies, obtained from data of the Hubble Deep Fields described in Sect. 9.2. Shown in blue, green and cyan are the luminosity functions at redshifts z ∼ 4, 5, and 6, respectively, whereas the magenta and red circles show the estimated luminosity functions at z ∼ 7 and 8. The curves show a Schechter-function fit to the data. Source: R.J. Bouwens et al. 2011, Ultraviolet Luminosity Functions from 132 z ∼ 7 and z ∼ 8 Lyman-break Galaxies in the Ultra-deep HUDF09 and Wide-area Early Release Science WFC3/IR Observations, ApJ 737, 90, p. 16, Fig. 12. ©AAS. Reproduced with permission
Going to even higher redshifts, the abundance of UV-selected galaxies decreases again, as can be seen in Fig. 9.41. The evolution is such that the characteristic luminosity decreases with higher z, and at the same time the faint-end slope steepens towards an estimated value of α ∼ −1. 8. Hence, for these very high redshifts, most of the luminosity in the UV is emitted from faint sources. Recent attempts to find credible z ∼ 10 galaxies using the Lyman-break technique yielded upper limits to the abundance of these objects, which yields upper limits to the luminosity function at this redshift. It appears that the decrease with z, visible in Fig. 9.41, accelerates towards even higher redshift.
In detail, these results are still burdened with quite some uncertainties, given the difficulties to identify very high redshift sources. Most of the conclusions are based on photometric redshifts only, since the spectroscopic verification of a z ∼ 7 galaxy is extremely difficult, given that all spectral features blueward of the Lyα emission line are invisible due to intergalactic absorption, and that the radiation redward of the Lyα-line is redshifted into the near-IR. Hence, some of the sources may be misidentified and are in fact lower-redshift objects. Furthermore, since the identification of these very high redshift sources requires very deep observations, carried out only in a small number of fields, one must be aware of sampling variance—the fact that the distribution in a single small field may not be representative of the overall distribution. However, the general trends just discussed are established by now, providing a clear view of the evolution of the galaxy population with cosmic time.
/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig42_HTML.jpg
Fig. 9.42
The comoving number density of galaxies with fixed rest-frame K-band luminosity, normalized by their current space density, as a function of redshift. The very different redshift dependence of high- and low-luminosity galaxies, in the sense that the abundance gets shifted towards lower luminosity—and thus lower stellar mass—objects with cosmic time, is called downsizing. Source: Cirasuolo et al. 2010, A new measurement of the evolving near-infrared galaxy luminosity function out to z = 4: a continuing challenge to theoretical models of galaxy formation, MNRAS 401, 1166, p. 1173, Fig. 7. Reproduced by permission of Oxford University Press on behalf of the Royal Astronomical Society
Optical/NIR luminosity function. The rest-frame optical light is a somewhat better indicator of the total stellar mass of galaxies than is the UV-radiation. However, only when going to the NIR is the luminosity of star-forming galaxies not dominated by the radiation from newly-born stars; in addition, the K-band light is rather unaffected by dust obscuration. To assess the rest-frame K-band emission of high-redshift galaxies, one needs mid-IR observations which became possible with the Spitzer Space Telescope. The results of a combined analysis of optical, near-IR and mid-IR data show again a dramatic change of the luminosity function with redshift. The characteristic density of galaxies Φ decreases with redshift, as one might expect—there should be fewer galaxies around at higher redshifts. For example, at z ∼ 2, Φ is about a factor 3.5 smaller than in the local Universe. In parallel to this, however, the characteristic luminosity L increases with z, by about one magnitude up to redshift 2. Hence it seems that a typical galaxy was brighter in the past. This phenomenon is quite counter-intuitive, given that the theory of structure formation predicts that more massive objects form in large abundance only at later redshifts—as follows from hierarchical structure formation. Another way to see this phenomenon is displayed in Fig. 9.42, which shows the comoving density of galaxies with fixed rest-frame K-band luminosity as a function of redshift, normalized to the corresponding local density. For rather low luminosities, the density decreases, but for high K-band luminosities, it increases by a factor ∼ 5 over a broad range in redshift, reaching a maximum at z ∼ 1. 5, and thereafter slowly decreases, but even at z ∼ 4 the density is still higher than in the local Universe.
It thus seems that the typical galaxy at high redshift has a larger stellar mass than currently, or that the ratio of high-mass to low-mass galaxies was substantially larger at high z. This implies that with increasing cosmic time, the galaxy population becomes increasingly dominated by those with lower mass. This phenomenon has received the name downsizing. Models of galaxy evolution in a hierarchical universe need to be able to describe this effect; we will return to this in Chap. 10.
A129044_2_En_9_Fig43_HTML.gif
Fig. 9.43
Estimates of the luminosity function in the infrared, for different redshift intervals. The total infrared luminosity was obtained by combining optical, near- and mid-IR data with models of the spectral energy distribution, and are shown as points connected with a thick blue curve, whereas the black curves show a fit by a parametrized Schechter-like function. These redshift-dependent luminosity functions are compared to the one at z = 0, indicated by the dashed green curve in each panel. Data points from different studies are included as different symbols. The vertical line displays an estimate of the completeness limit of the samples. Source: G. Rodighiero et al. 2010, Mid- and far-infrared luminosity functions and galaxy evolution from multiwavelength Spitzer observations up to z ∼ 2.5, A&A 515, A8, p. 17, Fig. 15. ©ESO. Reproduced with permission
/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig44_HTML.jpg
Fig. 9.44
From the luminosity functions shown in Fig. 9.43, the comoving total infrared luminosity density is obtained, and plotted as a function of redshift (left axis); on the right axis, the luminosity density is translated into an estimate of the corresponding star-formation rate density. Results from different studies are combined here, and they mutually agree quite well. The green and red dashed curves show the contribution to the luminosity density coming from galaxies with L IR ≥ 1011 L (LIRGs) and those with L IR ≥ 1012 L (ULIRGs). Source: G. Rodighiero et al. 2010, Mid- and far-infrared luminosity functions and galaxy evolution from multiwavelength Spitzer observations up to z ∼ 2. 5, A&A 515, A8, p. 18, Fig. 16. ©ESO. Reproduced with permission
Mid-IR luminosity function. Whereas the rest-frame UV-radiation indicates the level of star formation which is unobscured by dust, it misses those star-forming galaxies which are heavily obscured by dust. Their activity can best be seen in the rest-frame mid- and far-IR. At a fixed wavelength, the emitted flux depends on the amount of heat absorbed by the dust—and reradiated as thermal dust emission—and on the dust temperature. Therefore, the most reliable indicator of the obscured star formation rate is the bolometric infrared luminosity. Whereas this is not directly observable—the combination of sensitivity and field-of-view of far-IR detectors allows one to study only relatively bright objects—the combination of observations at optical, near-IR and mid-IR can be used to estimate the dust temperature and thus to derive the bolometric IR luminosity from the Spitzer 24 μm data and the derived dust temperature.
The corresponding evolution of the IR luminosity function is shown in Fig. 9.43 for several redshift bins. Although for the higher-redshift bins only the highest luminosity sources can be observed, the figure shows a dramatic evolution towards higher redshift: The number density of luminous sources increases by a large factor compared to the local one. The trend is similar to that shown in Fig. 9.42 for the K-band luminosity function, but even stronger. The increase of very strongly star-forming galaxies with redshift is stronger than that with large stellar masses. Combined with the evolution of largely unobscured star-formation, shown in Fig. 9.41, we therefore conclude that the star-forming activities had a much higher level in earlier epochs of cosmic evolution than it has today. We will come back to this point in more detail in Sect. 9.6
Integrating the Schechter function fits shown in Fig. 9.43 over luminosity, one obtains the total infrared luminosity emitted per unit comoving volume. The redshift evolution of this IR luminosity density is shown in Fig. 9.44. It must be pointed out that these estimates carry quite some uncertainty, since they require the extrapolation of the luminosity function to much fainter levels than those where data are available. In particular, the faint-end slope of the Schechter function is not at all well determined at high redshifts. Therefore, the detailed behavior of the luminosity density beyond z ∼ 1 may be slightly different from what is shown in the figure. However, the contribution of the LIRGs (defined as L ≥ 1011 L ) and ULIRGs (L ≥ 1012 L ) to the luminosity density, also shown in Fig. 9.44, is much better determined. While the IR luminosity density increases by a factor ∼ 20 between today and z ∼ 1, and stays roughly constant up to z ∼ 2. 5, the contribution from ULIRGs increases by at least a factor 100 over the same redshift range.
These results were confirmed with Herschel blank-field surveys, centered on fields for which multi-band observations were previously available (such as the GOODS fields or COSMOS). Observing in the far-IR, Herschel samples the peak of the spectral energy distribution directly, and fewer extrapolations are necessary to derive the bolometric infrared luminosity than required for using Spitzer data only. The analysis of the Herschel data showed that the evolution of the IR luminosity function with redshift is indeed dramatic. If one parametrizes the luminosity function as a Schechter function, the characteristic luminosity L in the infrared increases like (1 + z)3. 5 for 0 ≲ z ≲ 2, and ∝ (1 + z)1. 6 for 2 ≲ z ≲ 4. The normalization Φ of the Schechter function decreases with redshift, with  $$\varPhi ^{{\ast}} \propto (1 + z)^{-0.6}$$ for 0 ≲ z ≲ 1. 1, and  $$\propto (1 + z)^{-3.9}$$ for 1. 1 ≲ z ≲ 4. In agreement with the results shown in Fig. 9.43, the comoving space density of very IR-luminous sources increases by huge factors from today to higher redshifts, before it starts to decline beyond redshift z ∼ 3.
/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig45_HTML.jpg
Fig. 9.45
The restframe UB color, as a function of stellar mass, in six redshifts bins, as indicated in the six top panels. The two bottom panels show the color-stellar mass diagram for a proto-cluster at z = 1. 61, and the color-mass relation for all galaxies of the sample, irrespective of redshift. The straight line in each panel presents a fit to the red sequence, and the vertical dashed line indicates the completeness limit of the galaxy sample taken from the GMASS (Galaxy Mass Assembly ultradeep Spectroscopic Survey) project, in combination with multi-band photometry from optical to mid-infrared wavelengths in the GOODS-South field. The available high-resolution HST imaging allowed a morphological classification of the galaxies, according to which the symbols are color coded: early types (red), spirals (blue), irregulars (green), whereas cyan symbols are galaxies which are undetected in the optical bands and hence cannot be classified morphologically; those latter galaxies can appear only at the higher redshifts. The small inset in each panel shows the histogram of the color distribution. Source: P. Cassata et al. 2008, GMASS ultradeep spectroscopy of galaxies at z ∼ 2. III. The emergence of the color bimodality at z ∼ 2, A&A 483, L39, p. L40, Fig. 1. ©ESO. Reproduced with permission
/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig46_HTML.jpg
Fig. 9.46
Five z ∼ 2 galaxies from the CANDLES survey. Shown on the left is an IJH color composite of the galaxy, which corresponds to the restframe UV-to-optical wavelength range. The surface brightness in two filters centered on 0.85 and 1. 6 μm is shown in the next two columns. The fourth column displays the estimated UV restframe color, and the right column is the estimated stellar surface mass density. Source: S. Wuyts et al. 2012, Smooth(er) Stellar Mass Maps in CANDELS: Constraints on the Longevity of Clumps in High-redshift Star-forming Galaxies, ApJ 753, 114, p 6, Fig. 2. ©AAS. Reproduced with permission

9.4.2 The color-magnitude distribution

The color bimodality, seen prominently in the local population of galaxies (see Sect. 3.​1.​3), has been in place at least since z ∼ 2. As shown in Fig. 9.45, using spectroscopy of a 4. 5 μm-flux limited sample of galaxies in the GOODS South field, for which deep photometry is available over a wide range of optical and infrared bands (including HST and Spitzer), the color bimodality can be clearly seen in all redshift intervals.
At even higher redshift, the sample of galaxies on the red sequence gets increasingly contaminated by dusty star-forming galaxies. However, one can account for this reddening and obtain dust-corrected colors for those galaxies. After this correction, the color bimodality can be detected out to redshifts z ∼ 3, implying that already at young cosmic epochs, galaxies with an old stellar population coexisted with those which actively formed stars. Accordingly, the red sequence was formed early on, as can also be seen in Fig. 9.45.
This observational results implies that even at high redshifts, a large fraction of galaxies exists with a passively evolving stellar population. Whereas the star-forming galaxies—LBGs and SMGs—are the more spectacular objects at these high redshifts, many galaxies had formed their stars at even earlier epochs. From what we mentioned above—see, e.g., Fig. 9.42—the more massive galaxies seem to conclude most of the built-up of their stellar population at the highest redshifts. In parallel with the color-magnitude relation, also the local color-density relation was in place at least since z ∼ 1.
/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig47_HTML.jpg
Fig. 9.47
Left panel: The effective radius (i.e., the radius within which half the light is emitted) versus stellar mass of galaxies with Sérsic index n > 2. 5, representing early-type galaxies, for different redshifts. In each panel, the corresponding relation obtained in the local Universe is shown as solid curve with estimated uncertainties shown as error bars. Right panel: The mean effective radius (top) and velocity dispersion (bottom) as a function of redshift, for galaxies with stellar masses of M  ∼ 1011 M . The solid curve and grey band in the bottom panel shows two different models for the evolution of spheroidals. Source: Left: I. Trujillo et al. 2007, Strong size evolution of the most massive galaxies since z ∼ 2, MNRAS 382, 109, p. 115, Fig. 7. Reproduced by permission of Oxford University Press on behalf of the Royal Astronomical Society. Right: A.J. Cenarro & I. Trujillo 2009, Mild Velocity Dispersion Evolution of Spheroid-Like Massive Galaxies Since z ∼ 2, ApJ 696, L43, p. L46, Fig. 2. ©AAS. Reproduced with permission

9.4.3 The size and shape of high-redshift galaxies

The fact that the population of galaxies evolves strongly with redshift raises the expectation that the galaxies at high redshift are different from those in the local Universe. Here we will point out some of these differences.
The Hubble sequence and galaxy morphology. The majority of present day massive galaxies fall onto the morphological Hubble sequence (Fig. 3.​2). In addition, we have seen that local galaxies can be classified according to their color, with most of them being either member of the red sequence or the blue cloud, as seen in Fig. 3.​38, with a tight correspondence between the Hubble classification and the galaxy color and morphological parameters, such as the Sérsic index.
The situation at high redshifts is quite different. At z ≳ 3, most of the galaxies are strongly star forming, with a rather small population of quiescent galaxies. The star-forming objects do not appear at all to have a regular morphology, rather, they are irregular, or clumpy. In Fig. 9.46, five z ∼ 2 galaxies are shown as a IJH-color composite image. The irregular, knotty structure is easily seen, and many of the bright knots are clearly well separated from the center of the galaxy—these galaxies do not appear to fall on the Hubble sequence. These clumps have a characteristic size of ∼ 1 kpc, and they seem to be projected onto a kind of disk galaxy. In fact, using high angular resolution integral field spectroscopy, the rotation of several of such z ∼ 2 galaxies could clearly be shown. But these disk are not rotating quietly, in contrast to local disk galaxies, they have a very large velocity dispersion. This fact renders the interpretation of the observed velocity field in terms of Kepler rotation more complicted than for thin, kinematically cold disk galaxies.
However, we should keep in mind that the rest-frame UV light distribution is dominated by star-forming regions. The second and third column in Fig. 9.46 show the surface brightness of these galaxies in two different filters separately, and an estimate of the rest-frame UV color is shown in the fourth column. We can see that the clumps are typically significantly bluer than the rest of the galaxy, and hence their stellar population has a younger age than the underlying disk. The stellar mass contained in the clumps make a 7 % contribution to the total stellar mass of the galaxy, but they contribute about 20 % to the star-formation rate. Finally, the right column shows the reconstructed stellar surface mass density. Now the picture is a quite different one: the stellar mass is seen to be centrally concentrated, and no prominent off-center clumps are present.
Quiescent galaxies become more abundant towards lower redshift; it is estimated that the stellar mass contained in quiescent objects increases by a factor ∼ 15 between redshifts 3 and 1, and by another factor of ∼ 3 from there until today. In other word, the number of passive red galaxies has at least doubled since z = 1 until today, so that many of the early-type galaxies in the current Universe arrived on the red sequence at rather low redshift. For z > 2, peculiar galaxies dominate the galaxy population, with some quiescent, spheroidal galaxies already present then, but a negligible disk population. At a redshift around z ∼ 2, the abundance of spheroidal and disk galaxies together start to overtake the peculiar population, where this redshift depends on mass: at higher mass, the fraction of Hubble sequence-like galaxies is higher than at lower masses, indicating that they finish their morphological evolution earlier. Thus, starting from z ∼ 2, the Hubble sequence is gradually built up.
Size evolution. Red, quiescent galaxies at z ∼ 2 not only have a regular morphology compared to the clumpy star-forming galaxies, but they also are more massive and more compact. The latter aspects can be seen in Fig. 9.47, where in the left panel the effective radius is plotted as a function of stellar mass, for galaxies with Sérsic index n > 2. 5 and different redshift bins. Compared to the local population of early-type galaxies, higher-redshift spheroidal galaxies are significantly smaller at fixed stellar mass. The effective radius as a function of redshift, for a fixed stellar mass, is shown in the upper panel on the right. The size evolution is fitted with a power law of the form  $$r_{\mathrm{e}} \propto (1 + z)^{-1.48}$$ . The decrease in radius at fixed stellar mass by a factor of ∼ 3 at z ∼ 1. 5 implies that these galaxies have a stellar density larger by a factor ∼ 30 than present day early-type galaxies—these high-redshift galaxies are very different from the current population. The higher density also implies a larger velocity dispersion, which is indeed observed, as seen in the bottom panel on the right in Fig. 9.47.
Indeed, such compact galaxies are very rare in the local Universe—that means that the typical z ∼ 2 quiescent galaxy must have evolved significantly to fit into the local zoo of galaxies. Two principal possibility for this evolution exist: either the galaxies grow in size, at fixed stellar mass, or they accumulate more mass in their outer parts, thereby growing in mass and in size, such that they become less compact in this evolution. The latter possibility seems to be closer to the truth, as shown by simulations. Minor merging processes can yield an evolution that is compatible with the observational finding. Thus, early-type galaxies seem to grow inside-out: Their inner region was in place at earlier epochs, their outer parts were added lateron by merging processes.

9.4.4 The interstellar medium

The interstellar medium in high-redshift galaxies differs from that of local galaxies in a number of properties, of which we mention just a few here.
Metallicity. For local galaxies, there is a clear trend of increasing metallicity with increasing mass, as shown in Fig. 3.​40. A similar trend is observed for Lyman-break galaxies at z ∼ 2, except that the normalization of the mass-metallicity relation is smaller by a factor ∼ 2, as can be seen in Fig. 9.48. Of course, this result does not come unexpectedly, since at earlier redshifts, there was less time to enrich the ISM.
/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig48_HTML.jpg
Fig. 9.48
The metallicity of UV-selected z ∼ 2 galaxies, as a function of stellar mass (big dots with error bars). For comparison, the mass-metallicity relation for local SDSS galaxies is shown, as in Fig. 3.​40. Whereas the shape of the mass-metallicity relation is similar at both redshifts, the normalization is lower by about a factor of two at z ∼ 2. Source: D.K. Erb et al. 2006, The Mass-Metallicity Relation at z ≳ 2, ApJ 644, 813, p. 819, Fig. 3. ©AAS. Reproduced with permission
Whereas the metallicity of star-forming galaxies is lower than at the current epoch, at least some galaxies managed to enrich their ISM to about Solar values, as can be inferred from the metallicity of the broad-line emitting gas in high-redshifts QSOs. Not only did these objects form supermassive black holes with M  ≳ 109 M , but the chemical evolution in these objects was already mature at a small cosmic age. Of course, luminous high-z QSOs are rare and most likely populate the most massive halos available at these epochs.
Gas content. Second, the high star-formation rate implies the presence of a large reservoir of gas. As we have seen, in the current Universe the gas-mass fraction of even late-type spiral galaxies is below ∼ 30 %; in contrast to that, high-z star-forming galaxies typically have a gas-mass fraction of ∼ 50 %.
Instead of considering the absolute star-formation rate  $$\dot{M}$$ , it is often meaningful to study the specific star-formation rate,  $$\dot{M}/M_{{\ast}}$$ , i.e., the star-formation rate per unit stellar mass. This quantity has the units of an inverse time: the inverse of the specific star-formation rate is the time it would take to built up the stellar mass present if the star-formation rate would be a constant. At a fixed gas-mass fraction, this time is very similar to the time-scale on which all the gas in a galaxy is transformed into stars. The specific star-formation rate of Lyman-break galaxies increases by a factor ∼ 10 between the current epoch and z ∼ 1. 5, but then appears to stay remarkably constant out to z ∼ 7 (where, of course, the results at the highest redshifts carry an appreciable uncertainty).
Dust. The fact that the far-IR emission is the best indicator for star formation already indicates that these galaxies must contain a significant dust abundance. The dust temperature, which determines the spectral shape in the far-IR, can vary in these dusty galaxies over quite a substantial range, 25 K ≲ T d ≲ 65 K, as determined recently from Herschel observations. The impact of the dust temperature on the spectral shape means that single-band selection, e.g., the flux at 850 μm, can bias against objects with hotter dust temperature.
Whereas most of the properties of z ∼ 6 QSOs are indistinguishable from those of low-redshift QSOs, there are signs that they differ in their near-IR properties. In local QSOs, the UV/optical continuum emission is believed to be mainly due to the accretion disk, whereas the near-IR radiation is due to hot dust, heated by the AGN. The ratio of NIR-to-optical luminosity of z ≲ 5 QSOs is confined to a rather small range around ∼ 1. In a sample of 21 z ∼ 6 QSOs, there are two sources without detected NIR emission, yielding an upper limit to the NIR-to-optical flux ratio that is one order of magnitude smaller than the value typically observed. Using a control sample of more than 200 lower-z QSOs, not a single one has this flux ratio as low as the sources at z ∼ 6. The lack of detectable NIR emission can be attributed to the lack of dust in these systems.
A clue for the origin of this lack of dust is obtained from a second finding: the NIR-to-optical flux ratio for lower-z QSOs shows no correlation with the SMBH mass as estimated from the width of broad emission lines. In contrast to that, there seems to be a strong dependence of this flux ratio on the SMBH mass in the sample of z ∼ 6 QSO, in that the ratio increases with increasing M . A simple interpretation of this result could be that these high-redshift QSOs were able to build up their SMBH and the corresponding accretion disk, but that they were unable yet to form large masses of dust. The larger M , the more evolved is the AGN, and the more dust was created. It remains to be seen whether this interpretation survives further observational tests.

9.5 Background radiation at smaller wavelengths

The cosmic microwave background (CMB) is a remnant of the early hot phase of the Universe, namely thermal radiation from the time before recombination. As we extensively discussed in Sect. 8.​6, the CMB contains a great deal of information about our Universe. Therefore, one might ask whether background radiation also exists in other wavebands, which then might be of similar value for cosmology. The neutrino background that should be present as a relic from the early epochs of the Universe, in the form of a thermal distribution of all three neutrino families with T ≈ 1. 9 K (see Sect. 4.​4.​3), is likely to remain undiscovered for quite some time due to the very small cross section of these low-energy neutrinos.
Indeed, apparently isotropic radiation has been found in wavelength domains other than the microwave regime (Fig. 9.49). In this figure, the background radiation measured as ν I ν is plotted against wavelength, so that the curve shows the intensity per logarithmic frequency interval. Following the terminology of the CMB, these are called background radiation as well. However, the name should not imply that it is a background radiation of cosmological origin, in the same sense as the CMB. From the thermal cosmic history (see Sect. 4.​4), no optical or X-ray radiation is expected from the early phases of the Universe. Hence, for a long time it was unknown what the origin of these different background components may be.
/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig49_HTML.jpg
Fig. 9.49
Spectrum of cosmic background radiation, plotted as ν I ν versus wavelength. Besides the CMB, background radiation exists in the radio domain (cosmic radio background, CRB), in the infrared (CIB), in the optical/UV (CUVOB), in the X-ray (CXB), and at gamma-ray energies (CGB). With the exception of the CMB, all of these backgrounds can be understood as a superposition of the emission from discrete sources. Furthermore, this figure shows that the energy density in the CMB exceeds that of other radiation components, as was assumed when we considered the radiation density in the Universe in Chap. 4. Source: M.G. Hauser & E. Dwek 2001, The Cosmic Infrared Background: Measurements and Implications, ARA&A 39, 249, Fig. 1. Reprinted, with permission, from the Annual Review of Astronomy & Astrophysics, Volume 39 ©2001 by Annual Reviews www.​annualreviews.​org
At first, the early X-ray satellites discovered a background in the X-ray regime (cosmic X-ray background, CXB). Later, the COBE satellite detected an apparently isotropic radiation component in the FIR, the cosmic infrared background (CIB).
A129044_2_En_9_Fig50_HTML.gif
Fig. 9.50
The spectrum of the cosmic infrared background. The black arrows show the lower bound on the CIB, which is obtained from individually observed sources in these wavebands, i.e., by integrating the source counts at these frequencies up to the completeness limit. Blue arrows are lower bounds based on an extrapolation of the source counts to lower fluxes. The magenta arrows are upper limits, obtained from the Voyager spacecraft. The filled circles and stars with error bars are estimates, obtained with HST and several NIR instruments, as indicated. The solid magenta line presents an upper limit, obtained from the transparency of the Universe for very high-energy gamma rays with respect to electron-positron pair production (see Sect. 9.5.2); note that this upper limit is very close to the lower bound obtained from source counts. The open squares are estimates of the CIB from IRAS and the DIRBE instrument onboard COBE, while the solid curve at the longest wavelengths is the spectral estimate from FIRAS. Source: H. Dole et al. 2006, The cosmic infrared background resolved by Spitzer. Contributions of mid-infrared galaxies to the far-infrared background, A&A 451, 417, p. 426, Fig. 11. ©ESO. Reproduced with permission
In the present context, we simply denote the flux in a specific frequency domain, averaged over sky position at high Galactic latitudes, as background radiation. Thus, when talking about an optical background here, we refer to the sum of the radiation of all galaxies and AGNs per solid angle. The interpretation of such a background radiation depends on the sensitivity and the angular resolution of the telescopes used. Imagine, for instance, observing the sky with an optical camera that has an angular resolution of only one arcminute. A relatively isotropic radiation would then be visible at most positions in the sky, featuring only some very bright or very large sources. Thus, the background can be decomposed into a ‘resolved’ component, which can be attributed to individually identified sources, and the unresolved component. On improving the angular resolution, more and more individual sources become visible, so that a larger fraction of the background radiation is resolved. At optical wavebands, the Hubble Deep Fields have resolved essentially all of the background into individual sources. In analogy to this, one may wonder whether the CXB or the CIB can likewise be understood as a superposition of radiation from discrete sources.

9.5.1 The IR background

The first point to note from Fig. 9.49 is the relatively flat energy distribution between the UV- and the mm-regime. Since both, the UV-radiation and the far-IR radiation originate almost entirely from star-formation, the flat energy distribution implies that essentially half of the energetic photons emitted from newly-formed stars are absorbed by dust and reradiated in the FIR. Hence, estimates of the star-formation activity from UV-flux alone will on average be biased low by ∼ 50 %.
Absolute measurements of the intensity of the background radiation are difficult to obtain, since it requires an absolute calibration of the instruments. The filled circles and stars with error bars in Fig. 9.50 show estimates of the background radiation level in the optical and near-IR. The black arrows show the integrated light of all sources which are detectable in the deepest observations with HST; within the error bars of the estimated level of the background radiation, these results are compatible with the background being solely due to the superposition of individual optical and near-IR sources, i.e., galaxies and (to a lesser degree) AGNs.
Observations of background radiation in the infrared are very difficult to accomplish, in particular due to the thermal radiation from the instruments and the zodiacal light. However, the DIRBE and FIRAS instruments onboard COBE provided a measurement (in fact, the detection) of the CIB. The question now is whether the CIB can be understood as well as being due solely to individual sources.
Confusion limit. Since mid- and far-IR observations are only possible from space, finding the answer to that question is challenging. Infrared observatories in space have a rather small aperture which, together with the long wavelength, yields a rather large point-spread function (PSF). This implies that when one observes to low flux limits, where the mean angular separation of sources on the sky becomes comparable to the size of the PSF, these sources can not be separated. This yields a lower flux limit for the detection of individual sources, called the confusion limit. The smaller the telescope, the shallower is the confusion limit reached. For example, the flux limit down to which individual sources could be identified with the Spitzer satellite at 160 μm corresponds to only 7 % of the CIB at this wavelength. The much larger mirror on the Herschel satellite lowered the confusion limit such that individual sources can be identified which account for about 52 % of the CIB. Going to larger wavelength, the confusion limit is even more severe.
Stacking. However, one can dig deeper into the source counts with a technique called stacking. Taking the position of sources detected at some smaller wavelength (where the confusion limit is fainter), and adding up the flux in the longer wavelength band around all these positions, one obtains the mean long-wavelength flux of these sources. With this method, one will miss all fainter sources which do not have a detected counterpart in the short-wavelength input catalog, so that wavelength should be selected carefully. Given the characteristic spectrum of FIR-bright sources shown in Fig. 9.30, one expects that most of the sources radiating in the FIR will have an appreciable flux at 24 μm. Since Spitzer was particularly sensitive at this wavelength, the corresponding source catalog is best for a stacking analysis. Furthermore, if the redshifts of the sources selected at 24 μm is known, the stacking analysis can be used to determine the redshift distribution of the contributions to the CIB in the FIR. With stacking, the source counts can be followed to about three times lower flux than the confusion limit of individual sources permits.
The state of the art is defined by deep fields observed with the Herschel observatory, owing to its large aperture and sensitive instrumentation. From observing the well-studied GOODS, COSMOS, and ECDFS fields, where detailed multi-waveband data from other observatories are available, Herschel was able to attribute between 65 and 89 %, depending on wavelength, of the estimated CIB level between 100 and 500 μm to resolved sources or sources seen at 24 μm. A moderate extrapolation of the source counts to fainter flux limits then shows that the bulk, if not all, of the CIB comes from galaxies or AGNs.
In addition, the redshift distribution of the CIB could be determined. At wavelengths below ∼ 160 μm, more than half of the CIB radiation comes from sources at z < 1, whereas at longer wavelength, the source distribution shifts to increasingly higher redshifts. The major fraction of the CIB is due to galaxies with infrared luminosities in the range 1011 to 1012 L , i.e., due to LIRGs.

9.5.2 Limits on the extragalactic background light from γ-ray blazars

The added flux of sources, either individually detected or obtained from a stacking analysis, yields a lower limit to the extragalactic background light, which in the UV, optical and near-IR regime is smaller than estimates of the total intensity of the background, as seen in Fig. 9.50, although these latter measurements have fairly larger error bars. Hence, the question arises whether there are other sources of the background light not identified as individual sources—for example, very low surface brightness galaxies that could escape detection. In fact, this question can be answered from observations of blazars (see Sect. 5.​2.​6) at energies in the GeV and TeV regime, as will be described next.
Attenuation of γ -rays: Condition for pair production. High-energy photons from distant sources propagate through the extragalactic background radiation field. If the photon energy is high enough, then by colliding with one of the background photons, it may produce an  $$\mathrm{e}^{+}\mathrm{e}^{-}$$ -pair, in which case it does not reach the Earth. Thus, this pair production attenuates the flux from the source.
In order for pair production to occur, the product of the energies of the background-light photon (ε) and the photon from the source (E γ ) must be sufficiently high. If the two photons propagate in opposite direction (head-on collision), then the threshold condition is εE γ  > (m e c 2)2. In general, if the photon directions enclose an angle θ, this gets modified to
 $$\displaystyle{ \epsilon \,E_{\gamma } > \frac{2(m_{\mathrm{e}}c^{2})^{2}} {1-\cos \theta } \;. }$$
(9.3)
We see that the threshold energy is smallest for head-on collisions, where θ = π. The cross-section for this process is small for photon energies very close to the threshold, reaches its maximum at about twice the threshold energy (9.3), and decreases again for larger energies. At the maximum of the cross-section, the relation between the two photon energies can be written in practical units,
 $$\displaystyle{ \left ( \frac{E_{\gamma }} {1\,\mathrm{TeV}}\right ) = \frac{0.86} {1-\cos \theta }\,\left ( \frac{\lambda } {1\,\upmu \mathrm{m}}\right )\;, }$$
(9.4)
from which we see that for the attenuation of TeV photons, extragalactic background photons in the near-IR are most efficient, whereas radiation in the tens-of-GeV-regime can be attenuated by UV-photons.
A129044_2_En_9_Fig51_HTML.gif
Fig. 9.51
Top left: Model of the extragalactic background light, at three different redshifts. The high-amplitude peak at long wavelengths is the CMB and thus shows the evolution of the Planck spectrum with redshift. Top right: The proper photon number density per logarithmic energy interval, for the same three redshifts. Bottom left: The optical depth τ γ γ (E γ , z) for pair production, as a function of the γ-ray energy. Bottom right: The attenuation factor exp[−τ γ γ (E γ , z)] as a function of the γ-ray energy. Source: E. Dwek & F. Krennrich 2012, The Extragalactic Background Light and the Gamma-ray Opacity of the Universe, arXiv:1209.4661, Fig. 12. Reproduced by permission of the author
Optical depth. In order to derive the efficiency of the attenuation, one needs to calculate the optical depth τ γ γ (E γ , z), which depends on the energy of the γ-ray and the source redshift. To obtain τ γ γ (E γ , z), the pair-production cross section needs to be integrated along the line-of-sight to the source, multiplied by the spectral energy density of the background radiation; for this, the redshift evolution of the background radiation needs to be accounted for. Since the extragalactic background light can be observed only at the current redshift, one needs to model its redshift evolution, based on what is known about the source population. A particular model is shown in the top left panel of Fig. 9.51, which also includes the CMB, and the corresponding photon number density per logarithmic photon energy interval is shown in the top right panel. Based on the background light model, the optical depth for pair production can be calculated, which is shown in the bottom left panel of the same figure. τ γ γ (E γ , z) is a strong function of both, the γ-ray energy and the redshift. The plateau in τ γ γ at energies ∼ 2 TeV is due to the minimum of the background radiation spectrum at ∼ 10 μm.
The observed flux S obs(E γ ) is related to the intrinsic (i.e., non-attenuated) flux S int(E γ ) by
 $$\displaystyle{ S_{\mathrm{obs}}(E_{\gamma }) = S_{\mathrm{int}}(E_{\gamma })\,\exp [-\tau _{\gamma \gamma }(E_{\gamma },z)]\;, }$$
(9.5)
where the attenuation factor  $$\mathrm{e}^{-\tau _{\gamma \gamma }}$$ is plotted in the bottom right panel of Fig. 9.51. We can see that the attenuation factor has a very steep decline with photon energy; for example, based on this model we would not expect to see 20 TeV photons from any source at z ≳ 0. 1. This steep, exponential decline implies that the attenuation factor is very sensitive to the model of the extragalactic background light; conversely, if observations yield constraints on the attenuation factor, then strong constraints on the background light can be obtained, at wavelengths depending on the detection of the attenuation, according to (9.4).
Observational constraints on the attenuation. As mentioned in Sect. 5.​5.​4, blazars can emit at GeV and TeV energies, most likely caused by their jet pointing towards us. The Fermi Gamma-Ray Space Telescope, operating in the energy range between 200 MeV and 300 GeV, and the air Cherenkov observatories H.E.S.S., MAGIC and VERITAS, observing in the range between ∼ 50 GeV and 100 TeV, have detected more than 1000 blazars in the GeV-range and more than 30 at TeV energies. Whereas blazars in the GeV-range are observed out to redshifts z ≥ 1, essentially all the TeV blazars are at low redshift, most of them having z ≲ 0. 2. This is indeed what one expects, based on the results in Fig. 9.51.7 In principle, these observations of the spectral energy distribution could be used to determine the attenuation factor; however, in order to employ (9.5), one needs some knowledge about the intrinsic flux distribution S int(E γ ).
There are various ways how realistic estimates for τ γ γ can be obtained from the observations. The first of these is to base the intrinsic flux distribution on models of the γ-ray emission, and constrain these models by observations at somewhat lower photon energies. However, the models are sufficiently uncertain to preclude very accurate predictions, and thus the corresponding results on τ γ γ are correspondingly uncertain. Second, since the relativistic electrons responsible for the inverse Compton effect that presumably causes the γ-ray emission, are expected to result from acceleration by shock fronts, as mentioned in Sect. 5.​1.​3, the slope of the electron distribution can not be arbitrarily flat, and thus the resulting inverse Compton radiation is also limited in slope; in the notation of Sect. 5.​1.​3, s ≳ 2 and thus α ≳ 0. 5. Assuming this value of the spectral index as a limit for the intrinsic flux distribution, the observed energy distribution can be translated into upper bounds on the attenuation. An even weaker assumption is used by a third methods, where one requires that the intrinsic flux S int(E γ ) = S obs(E γ ) exp[τ γ γ (E γ , z)] does not (exponentially) increase with photon energy, as would be the case if the background light intensity would be overestimated.
Results. Depending on which of these methods are used, the results will differ slightly. A particular result is shown in Fig. 9.50, where the magenta curve indicates the upper bound on the extragalactic background light obtained from the high-energy observations of blazars. This upper bound is almost coincident with the lower bound obtained from the resolved source counts in the UV, optical and near-IR regime, strongly arguing that there are no other significant contributions of the background light than the observed galaxies and AGN. It thus seems that the spectral intensity of the background light in this spectral regime is now rather well determined. This conversely implies that the optical depth τ γ γ is very well constrained, which in turn allows us to derive the intrinsic flux distribution from the observed one. In the future, this method will therefore yield more detailed constraints on the emission mechanism for high-energy radiation from blazars and other AGNs.
/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig52_HTML.jpg
Fig. 9.52
Measurement of the cosmic X-ray background over a wide range of photon energies, measured by different satellites and instruments. Source: R. Gilli 2013, The cosmic X-ray background: abundance and evolution of hidden black holes, arXiv:1304.3665, Fig. 1. Reproduced by permission of the author

9.5.3 The X-ray background

The first X-ray experiment in astronomy, a balloon flight in 1962, discovered a diffuse X-ray emission across the sky, confirmed by the first X-ray satellites which discovered not only a number of extragalactic X-ray sources (such as AGNs and clusters of galaxies), but also an apparently isotropic radiation component. The spectrum of the cosmic X-ray background (CXB) is a very hard (i.e., flat) power law, cut off at an energy above ∼ 40 keV, which can roughly be described by
 $$\displaystyle{ I_{\nu } \propto E^{-0.3}\,\exp \left (-\frac{E} {E_{0}}\right )\;, }$$
(9.6)
with E 0 ∼ 40 keV. A recent estimate of the spectrum of the CXB is shown in Fig. 9.52. The estimates from different instruments agree in general, though differences in the level are clearly visible. These differences can have a number of origins, including cosmic variance (the spectral shape of the CXB is usually determined from rather small fields, so there could be variations from field to field), stray light entering the telescope, and remaining calibration uncertainties of the instruments. Together, the CXB is known with an uncertainty of ∼ 20 %.
Initially, the origin of this radiation was unknown, since its spectral shape was different from the spectra of sources that were known at that time. For example, it was not possible to obtain this spectrum by a superposition of the spectra of known AGNs.
/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig53_HTML.jpg
Fig. 9.53
In the left panel, the total intensity of discrete sources with an individual flux > S in the energy range 2 keV ≤ E ≤ 10 keV is plotted (thick curve), together with the uncertainty range (between the two thin curves). Most of the data are from a 3 × 105 s exposure of the Chandra Deep Field. The dashed lines show different measurements of the CXB flux in this energy range; depending on which of these values is the correct one, between 60 and 90 % of the CXB in the Chandra Deep Field at this energy is resolved into discrete sources. In the right panel, the hardness ratio HR—specifying the ratio of photons in the energy range 2 keV ≤ E ≤ 10 keV to those in 0. 5 keV ≤ E ≤ 2 keV,  $$\mathrm{HR} = (S_{>2\mathrm{keV}} - S_{<2\mathrm{keV}})/(S_{>2\mathrm{keV}} + S_{<2\mathrm{keV}})$$ —is plotted as a function of redshift, for 84 sources in the Chandra Deep Field with measured redshifts. This plot indicates that the HR decreases with redshift; this trend is expected if the X-ray spectrum of the AGNs is affected by intrinsic absorption. The dashed curves show the expected value of HR for a source with an intrinsic power-law spectrum I ν  ∝ ν −0. 7, which is observed through an absorbing layer with a hydrogen column density of N H, by which these curves are labeled. Since low-energy photons are more strongly absorbed by the photoelectric effect than high-energy ones, the absorption causes the spectrum to become harder, thus flatter, at relatively low X-ray energies. This implies an increase of the HR (also see the bottom panel of Fig. 6.​22). This effect is smaller for higher redshift sources, since the photon energy at emission is then larger by a factor of (1 + z). Source: P. Tozzi et al. 2001, New Results from the X-Ray and Optical Survey of the Chandra Deep Field-South: The 300 Kilosecond Exposure. II., ApJ 562, 42, p. 48, 49, Figs. 7, 9. ©AAS. Reproduced with permission
ROSAT, with its substantially improved angular resolution compared to earlier satellites (such as the Einstein observatory), conducted source counts at much lower fluxes, based on some very deep images. From this, it was shown that at least 80 % of the CXB in the energy range between 0.5 and 2 keV is emitted by discrete sources, of which the majority are AGNs. Hence it is natural to assume that the total CXB at these low X-ray energies originates from discrete sources, and observations by XMM-Newton and Chandra have confirmed this.
However, the X-ray spectrum of normal AGNs is different from (9.6), namely it is considerably steeper (about S ν  ∝ ν −0. 7). Therefore, if these AGNs contribute the major part of the CXB at low energies, the CXB at higher energies cannot possibly be produced by the same AGNs. Subtracting the spectral energy of the AGNs found by ROSAT from the CXB spectrum (9.6), one obtains an even harder spectrum, resembling very closely that of thermal bremsstrahlung. Therefore, it was supposed for a long time that the CXB is, at higher energies, produced by a hot intergalactic gas at temperatures of k B T ∼ 30 keV.
This model was excluded, however, by the precise measurement of the thermal spectrum of the CMB by COBE, showing that the CMB has a perfect blackbody spectrum. If a postulated hot intergalactic gas were able to produce the CXB, it would cause significant deviations of the CMB from the Planck spectrum, namely by the inverse Compton effect (the same effect that causes the SZ effect in clusters of galaxies—see Sect. 6.​4.​4). Thus, the COBE results clearly ruled out this possibility.
Deep observations with Chandra and XMM (e.g., in the CDFS shown in Fig. 9.14) have finally resolved most of the CXB also at higher energies, as seen in Fig. 9.53. From source counts performed in such fields, more than 75 % of the CXB in the energy range of 2 keV ≤ E ≤ 10 keV could be resolved into discrete sources. Again, most of these sources are AGNs, but typically with a significantly harder (i.e., flatter) spectrum than the AGNs that are producing the low-energy CXB. Such a flat X-ray spectrum can be produced by photoelectric absorption of an intrinsically steep power-law spectrum, where photons closer to the ionization energy are more efficiently absorbed than those at higher energy. According to the classification scheme of AGNs discussed in Sect. 5.​5, these are Type 2 AGNs, thus Seyfert 2 galaxies and QSOs with strong intrinsic self-absorption. We should recall that Type 2 QSOs have only been detected by Chandra—hence, it is no coincidence that the same satellite has also been able to resolve the high-energy CXB.
However, at even higher energies most of the CXB was still unaccounted for—even the observed Type-2 AGNs could not account for it. It thus seems that there is a population of sources in the Universe which dominate the X-ray emission at high energies, still escape the observations at low X-ray frequencies. These could be heavily obscured AGNs, where only the hard X-rays manage to escape the emitting region. With the X-ray telescope onboard the Swift satellite, a significant number of such heavily obscured AGNs were found. Their estimated number density, together with their spectral energy distribution, make it plausible that they are the missing population of ‘hidden black holes’ responsible for the hard CXB.

9.6 The cosmic star-formation history

The evolution of the galaxy luminosity function with redshift, discussed in Sect. 9.4.1, provides clear evidence that the galaxy population is strongly changing with cosmic epoch. In particular, the z-dependence of the luminosity functions in the UV and the IR shows that the rate at which new stars form in the Universe must be a function of time—such that the average star-formation rate (SFR) was considerably larger at high redshifts. In this section, we consider this evolution of the SFR, together with the evolution of the stellar density. Of course, these two quantities are related: The stellar density at redshift z is the integral of the SFR per unit volume over time, from the earliest epochs to the one corresponding to z. The combination of sensitive space observatories with large ground-based telescopes equipped with modern instruments allows us to trace the SFR up to very high redshifts.
We define the star-formation rate (SFR) as the mass of the stars that form per unit time in a galaxy, typically given in units of M ∕yr. For our Milky Way, we find a SFR of ∼ 3M ∕yr. Furthermore, we define the star-formation rate density as the mass of stars that are formed per unit time and per unit (comoving) volume, expressed in units of  $$M_{\odot }\,\mathrm{yr}^{-1}\,\mathrm{Mpc}^{-3}$$ .
The importance of the initial mass function. Since the observable signatures for star formation are obtained only from massive stars, their formation rate needs to be extrapolated to lower masses to obtain the full SFR, by assuming an IMF (initial mass function; see Sect. 3.​5.4). Typically, a Salpeter-IMF is chosen between 0. 1M  ≤ M ≤ 100M . However, there are clear indications that the IMF may be flatter for M ≲ 1M than described by the Salpeter law, and several descriptions for such modified IMFs have been developed over the years, mainly based on observations and interpretation of star-forming regions in our Milky Way or in nearby galaxies. The total stellar mass, obtained by integration over the IMF, is up to a factor of ∼ 2 lower in these modified IMFs than for the Salpeter IMF. Thus, this factor provides a characteristic uncertainty in the determination of the SFR from observations; a similar, though somewhat smaller uncertainty applies to the stellar mass density whose estimation also is mainly based on the more massive stars of a galaxy which dominate the luminosity. Furthermore, the IMF need not be universal, but may in principle vary between different environments, or depend on the metallicity of the gas from which stars are formed. Whereas there has not yet been unambiguous evidence for variations of the IMF, this possibility must always be taken into account.

9.6.1 Indicators of star formation

We will start by discussing the most important indicators of star formation.
Emission in the far infrared (FIR). This is radiation emitted by warm dust which is heated by hot young stars. Observations yield for the approximate relation between the FIR luminosity and the SFR
 $$\displaystyle{\frac{\mathrm{SFR_{FIR}}} {M_{\odot }/\mathrm{yr}} \sim \frac{L_{\mathrm{FIR}}} {5.8 \times 10^{9}L_{\odot }}\;.}$$
For this relation it is assumed that all the energetic photons from newly born hot stars are absorbed locally and heat the dust; more generally, this expression yields the SFR that is dust enshrouded. The wavelength range over which L FIR is determined should be large, covering the two decades from 8 μm to 1 mm, so that this luminosity is essentially independent of the dust temperature. However, in most cases the observation do not cover such a broad spectral region, so that interpolation and extrapolation of the spectral behavior is required to determine L FIR, based on template spectra constructed from very well observed sources. This procedure therefore carries an intrinsic uncertainty in the determination of the SFR. For large samples of sources, one often uses the 24 μm flux to obtain an estimate of L FIR; this wavelength is seen as a good compromise between the need to go to large wavelengths and the decreasing sensitivity and field-of-view of infrared instrumentation. The Spitzer Space Telescope played a very important role in the studies of star formation at these wavelengths.
Radio emission by galaxies. A very tight correlation exists between the radio luminosity of galaxies and their luminosity in the FIR, over many orders of magnitude of the corresponding luminosities. Since L FIR is a good indicator of the star-formation rate, this should apply for radiation in the radio as well (where we need to disregard the radio emission from a potential AGN component). The synchrotron radio emission of normal galaxies originates mainly from relativistic electrons accelerated in supernova remnants (SNRs). Since SNRs appear shortly after the beginning of star formation, caused by core-collapse supernovae at the end of the life of massive stars in a stellar population, radiation from SNRs is a nearly instantaneous indicator of the SFR. Once again from observations, one obtains
 $$\displaystyle{\frac{\mathrm{SFR_{1.4\,GHz}}} {M_{\odot }/\mathrm{yr}} \sim \frac{L_{\mathrm{1.4\,GHz}}} {8.4 \times 10^{27}\,\mathrm{erg\,s^{-1}\,Hz^{-1}}}\;.}$$
H α emission. This line emission comes mainly from the Hii-regions that form around young hot stars with M ≳ 10M . As an estimate of the SFR, one uses
 $$\displaystyle{ \frac{\mathrm{SFR}_{\mathrm{H}\alpha }} {M_{\odot }/\mathrm{yr}} \sim \frac{L_{\mathrm{H}\alpha }} {1.3 \times 10^{41}\,\mathrm{erg\,s^{-1}}}\;.}$$
For redshifts z ≳ 2, the observed Hα lines moves into the near-IR part of the spectrum and is thus much more difficult to observe. One can also employ other emission lines, such as Hβ or Lyα. However, whereas Hβ can be observed in the optical to higher redshifts, due to its shorter wavelength, it is a weaker line. For Lyα, the uncertainty to convert line flux into a SFR is much larger, since it is a transition to the ground state of hydrogen (a so-called resonance line). Because of that, a Lyα photon is easily absorbed by neutral hydrogen, which is then excited and reemits a Lyα photon in a random direction. Effectively, thus, this process is a scattering of the photon. In order to leave the interstellar medium of a galaxy, a Lyα photon may be subject to many such scatters, which implies that is path inside the ISM is very much enhanced. This longer path then also increases the probability for the photon to become absorbed by dust. Therefore, the conversion of Lyα flux that leaves a galaxy and the SFR is burdened with a large uncertainty. Alternatively, one can also consider recombination lines of hydrogen coming from transition between higher energy states of the atom, such a Bracket γ, a transition that occurs in the NIR of the spectrum, or transition lines that have millimeter or radio wavelengths.
UV radiation. This is mainly emitted by O and B stars, i.e., by hot young stars thus indicating the SFR in the most recent past, with
 $$\displaystyle{\frac{\mathrm{SFR_{UV}}} {M_{\odot }/\mathrm{yr}} \sim \frac{L_{\mathrm{UV}}} {7.2 \times 10^{27}\,\mathrm{erg\,s^{-1}\,Hz^{-1}}}\;.}$$
This relation assumes that the UV-flux can leave the galaxy without being attenuated by dust absorption (and it neglects that there may be an AGN contribution to the UV luminosity). However, in most galaxies this is not a reasonable assumption, and the observed L UV must be corrected for this effect. Due to the wavelength-dependence of dust absorption, extinction is always connected to reddening, thus affecting the spectral slope of the UV-radiation. One therefore expects that the redder the UV-spectrum, the larger are the effects of dust obscuration. To quantify this effect, one has to assume an intrinsic, unobscured spectral shape, and to make assumptions about the dust properties, specifically regarding its wavelength dependence of the extinction coefficient (see Fig. 2.​6). The unobscured spectral shape can be obtained from the shape of the IMF at the high-mass end (i.e., over that mass region of stars from which the UV-radiation is emitted), whereas there is considerable uncertainty about the variation of dust properties between different objects.
As a result of this procedure, one finds that only a small fraction of UV-photons actually leaves an UV-selected galaxy. A typical value for this escape fraction in a Lyman-break galaxy is ∼ 0. 2, which means that the observed UV-flux has to be corrected by a factor ∼ 5 to obtain the corresponding SFR.
X-ray luminosity. We have seen (Fig. 9.15) that non-active galaxies are X-ray emitters. Most of the X-ray emission is due to high-mass X-ray binaries which are members of a young stellar populations. About 25 % of the X-ray emission from a normal galaxy is due to bremsstrahlung from a hot interstellar medium; since its heating is provided by star-formation activity, it should also scale with the star-formation rate. Hence, if a contribution from an AGN can be excluded, the X-ray luminosity should be a good indicator for the star-formation rate. One finds a tight relation,
 $$\displaystyle{\frac{\mathrm{SFR_{X-ray}}} {M_{\odot }/\mathrm{yr}} \sim \frac{L_{\mathrm{X-ray}}} {3.5 \times 10^{39}\,\mathrm{erg\,s^{-1}}}\;,}$$
where the X-ray luminosity is integrated from 0. 5 to 8 keV, and where the scatter in this relation is estimated to be less than a factor of 1.5.
Comparison. Applied to individual galaxies, each of these estimates is quite uncertain, which can be seen by comparing the resulting estimates from the various methods (see Fig. 9.54). For instance, Hα and UV photons are readily absorbed by dust in the interstellar medium of the galaxy or in the star-formation regions themselves. Therefore, the relations above should be corrected for this self-absorption, which is possible when the reddening can be obtained from multi-color data. It is also expected that the larger the dust absorption, the stronger the FIR luminosity will be, causing deviations from the linear relation SFRFIR ∝ SFRUV. After the appropriate corrections, the values for the SFR derived from the various indicators are quite similar on average, but still have a relatively large scatter.
There are also a number of other indicators of star formation. The fine-structure line of singly ionized carbon at λ = 157. 7 μm is of particular importance as it is one of the brightest emission lines in galaxies, which can account for a fraction of up to 1 % of their total luminosity. The large abundance of carbon, together with the fact that this transition can be collisionally excited even at low temperature, result in this line to have a major role in the cooling of the neutral interstellar medium, and thus signifies the presence of star-forming regions. At its wavelength, this line is difficult to observe and until recently has been detected only in star-forming regions in our Galaxy and in other local galaxies. However, more recently this line was detected from the host galaxies of high-redshift QSOs and SMGs, where it shifted into more accessible spectral windows.
/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig54_HTML.jpg
Fig. 9.54
Correlations of the star formation rates in a sample of galaxies, as derived from observation in different wavebands. In all four diagrams, the dashed line marks the identity relation SFR i  = SFR j ; as is clearly seen, using the Hα luminosity and UV radiation as star-formation indicators seems to underestimate the SFR. Since radiation may be absorbed by dust at these wavelengths, and also since the amount of warm dust probably depends on the SFR itself, this effect can be corrected for, as shown by the solid curves in the four panels. Source: A.M. Hopkins et al. 2001, Toward a Resolution of the Discrepancy between Different Estimators of Star Formation Rate, AJ 122, 288, p. 291, Figs. 2, 3. ©AAS. Reproduced with permission

9.6.2 Redshift dependence of the star formation: The Madau diagram

The density of star formation, ρ SFR, is defined as the mass of newly formed stars per year per unit (comoving) volume, typically measured in  $$M_{\odot }\,\mathrm{yr}^{-1}\,\mathrm{Mpc}^{-3}$$ . Therefore, ρ SFR as a function of redshift specifies how many stars have formed at any time. By means of the star-formation density we can examine the question, for instance, of whether the formation of stars began only at relatively low redshifts, or whether the conditions in the early Universe were such that stars formed efficiently even at very early times.
Investigations of the SFR in galaxies, by means of the above indicators, and source counts of such star-forming galaxies, allow us to determine ρ SFR. The plot of these results (Fig. 9.55) is sometimes called “Madau diagram”. In about 1996, Piero Madau and his colleagues accomplished, for the first time, an estimate of the SFR at high redshifts from Lyman-break galaxies in the Hubble Deep Field North. For these early results, the intrinsic extinction was neglected. In order to correct for this extinction, the progress in FIR and sub-millimeter astronomy was extremely important, as we saw in Sect. 9.3.3.
There is a strong increase in ρ SFR from the current epoch to z = 1 by about a factor 10, a further slight increase towards z ∼ 2, and a decrease at redshifts beyond z ∼ 3. These results have more recently been confirmed by investigations with the Spitzer and Herschel satellites, observing a large sample of galaxies at FIR wavelengths. Whereas the star-formation rate density at low redshifts is dominated by galaxies which are not very prominent at FIR wavelength, this changes drastically for redshifts z ≳ 0. 7, above which most of the star-formation activity is hidden from the optical view by dust.8 The increasing importance of dust-obscured star formation is concluded from the very strong redshift evolution of the infrared luminosity function of galaxies shown in Fig. 9.43, and the corresponding evolution of the infrared luminosity density (Fig. 9.44).
Integrating the star-formation density over cosmic time, one obtains the stellar mass density as a function of redshift, shown in Fig. 9.56. From this we conclude that most stars in the present-day Universe were already formed at high redshift: star formation at earlier epochs was considerably more active than it is today. Although the redshift-integrated star-formation rate and the mass density of stars determined from galaxy surveys slightly deviate from each other, the degree of agreement is quite satisfactory if one recalls the assumptions that are involved in the determination of the two quantities: besides the uncertainties discussed above in the determination of the star-formation rate, we need to mention in particular the shape of the IMF of the newly formed stars for the determination of the stellar mass density. In fact, Fig. 9.56 shows that we have observed the formation of essentially the complete current stellar density.
The difference between the observed stellar mass density and the one predicted from Fig. 9.55 is largest at high redshifts, which may be due to the uncertainties with which ρ SFR is determined at high redshifts. With the more recent very deep fields observed with HST, the unobscured star-formation rate can be estimated at larger redshifts, based on the luminosity function of Lyman-break galaxies (Fig. 9.41). The result is shown in Fig. 9.57, which shows a steep decline of ρ SFR towards higher redshift.9 Hence, the bottom line is that there was an epoch in the Universe, between redshifts ∼ 1 and ∼ 4, where the star-formation activity was largest. This epoch coincides with the period in our Universe where the QSO activity was highest—indicating that the built-up of the supermassive black hole mass in the Universe happened in parallel to the formation of the stellar population. The close correlation between SMBH mass and properties of the stellar population discussed in Sect. 3.​8.​3 may thus find a first explanation in this parallel evolution.
/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig55_HTML.jpg
Fig. 9.55
The comoving star-formation density ρ SFR as a function of redshift, where the different symbols denote different indicators used for the determination of the star-formation rate. This plot shows the history of star formation in the Universe. Clearly visible is the strong decline for z < 1; towards higher redshifts, ρ SFR seems to remain approximately constant. The curve is an empirical fit to the data. Source: E.F. Bell 2004, Galaxy Assembly, astro-ph/0408023, Fig. 1. Reproduced by permission of the author
/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig56_HTML.jpg
Fig. 9.56
Redshift evolution of the mass density in stars, as measured from various galaxy surveys. The solid curve specifies the integrated star-formation density from Fig. 9.55. Source: E.F. Bell 2004, Galaxy Assembly, astro-ph/0408023, Fig. 2. Reproduced by permission of the author
Different modes of star formation. Whereas most of the star formation in the local Universe occurs in spiral and irregular galaxies at a modest rate (so-called quiescent star formation), the star-formation activity at higher redshifts was dominated by bursts of star formation, as evidenced in the sub-mm galaxies and in LBGs. At a redshift z ∼ 1, the latter has apparently ceased to dominate, yielding the strong decline of the star-formation rate density from then until today. This behavior may be expected if bursts of star formation are associated with the merging of galaxies ; the merger rate declines strongly with time in models of the Universe dominated by a cosmological constant. This transition may also be responsible for the onset of the Hubble sequence of galaxy morphologies around z ∼ 1.
The starburst-AGN connection. The just-mentioned coincidence of the ‘QSO epoch’ with the peak of the star-formation activity can either have a statistical origin, or there can be a connection object-by-object, in the sense that QSO are hosted in galaxies with active star formation. For physical reasons, one would in fact expect such a direct connection, since both processes, star formation and AGN fueling, need a gas supply.
A129044_2_En_9_Fig57_HTML.gif
Fig. 9.57
Based on deep HST imaging in the optical and near-infrared, the star-formation density ρ SFR and stellar mass density was estimated to higher redshifts. Left panel: Blue points (and right axis) show the UV-luminosity density as a function of redshift, which is proportional to the unobscured star-formation rate density. The red points (left axis) show the star-formation rate density where a correction for obscured star formation is included; this correction is assumed to be negligible for z ≳ 7. The contribution from galaxies with M UV ≤ −18 in AB magnitudes are added up for these estimates. Black and green points are estimates of ρ SFR and the UV-luminosity density for z ≤ 3. The points at z ∼ 10 is based on LBG candidates detected behind lensing clusters of the CLASH survey. Right panel: The stellar mass density as a function of redshift. For high redshifts, ρ decreases approximately  $$\propto (1 + z)^{-3.4}$$ , as indicated by the dashed grey curve. Source: Left: R.J. Bouwens et al. 2012, A Census of Star-Forming Galaxies in the z ∼ 9 − 10 Universe based on HST+Spitzer Observations Over 19 CLASH clusters: Three Candidate z ∼ 9 − 10 Galaxies and Improved Constraints on the Star Formation Rate Density at z ∼ 9. 2, arXiv:1211.2230, Fig. 10. Right: V. Gonzalez et al. 2011, Evolution of Galaxy Stellar Mass Functions, Mass Densities, and Mass-to-light Ratios from z ∼ 7 to z ∼ 4, ApJ 735, L34, p. 6, Fig. 4. ©AAS. Reproduced with permission
Indeed, in a large fraction of QSOs, clear signs of star-formation activity are found, in some cases at a level where the QSO host galaxy appears as a ULIRG. Conversely, in many of the low-redshift star-forming galaxies, signs of the presence of an AGN are seen. Furthermore, a direct connection between these two processes in a given galaxy does not necessarily have to be observable: it is conceivable that a fresh supply of gas (say, from a major merging of two galaxies) first leads to a strong star-formation activity, and that the accretion onto the central black hole occurs with some delay—or in reverse order.
Statistical studies based on the MIR and FIR emission properties of X-ray selected AGNs suggest that for low-luminosity AGNs, there is no correlation between AGN luminosity and star-formation rate. However, at high AGN luminosity, such a correlation is indeed found, providing a strong hint for the connection between the built-up of the black hole mass and the stellar population in individual sources.

9.6.3 Summary: High-redshift galaxies

In this chapter, we have considered various aspects of galaxies in the high-redshift Universe. Our discussion of this very quickly evolving field is not complete, but concentrates on some of the central issues. Before we move to another class of high-redshift sources, we want to summarize some of the points mentioned before:
  • High-redshift galaxies can be selected by a number of different methods, the most famous one being the Lyman-break technique, other multi-(optical and NIR) band selections, narrow-band imaging targeting highly-redshifted Lyα emitters, mid-infrared selection, and far-infrared/(sub-)millimeter selection. Spectroscopic confirmation of these candidates can be quite challenging, in particular for very dusty galaxies which can be very faint in the optical and NIR spectral regime, and when the source redshift approaches ∼ 7, so that the Lyα line is shifted out of the optical window.
  • As is true for other situations as well, the properties of the galaxy sample obtained depend on the selection method. A comparison of different samples can therefore be difficult, and must proceed with great care. Lyman-break galaxies at z ∼ 3 have a stellar mass smaller by a factor ∼ 10 than sub-millimeter galaxies, but larger masses than Lyman-alpha emitters.
  • The galaxy population at high redshift is distinctly different from that in the current Universe. Most galaxies at z > 2 do not fit into Hubble’s morphological classification, but show irregular light distributions. The star-formation activity in the Universe was far more intense in the past than it is now. At z ∼ 2. 5, some 10 % of all stars had been formed, and about 50 % of the stars in the local Universe were in place at z ∼ 1. Correspondingly, the average star-formation rate of distant galaxies is much higher than that of local galaxies. This is reflected in the strong evolution of the galaxy luminosity function in wavebands which strongly respond to the star-formation activity—most notably at mid-IR, far-IR and (sub-)millimeter wavelengths. Similarly, the star-formation rate density is a strongly evolving function of redshift, with a more than tenfold increase between today and z ∼ 1, an extended period of redshift lasting to z ∼ 3 or 4, where the star-formation density stays at a high rate, before declining towards even higher redshifts.
  • On the other hand, even at z ∼ 2. 5, about half of the most massive galaxies are quiescent, that is, they must have formed their large stellar population at even higher redshifts. From the evolution of the luminosity function with redshift, it appears that the most massive galaxies formed most of their stars early on, and lower-mass galaxies finish most of their evolution at lower redshifts. This trend has been termed ‘downsizing’ in the literature.
  • The mean metallicity of galaxies evolves with redshift. At a fixed stellar mass, the metallicity of galaxies at z = 2 is about smaller by a factor ∼ 2 than today, and a further factor of ∼ 2 decrease is found at z ∼ 3. 5. On the other hand, the gas of high-redshift QSOs seems to be fairly enriched with metals, approaching Solar metallicity. The dust content of galaxies appears to decrease towards the highest available redshifts, with dust-poor and almost dust-free QSOs detected at z ∼ 6.
  • Except for the CMB, which is a relic of the Big Bang, the radiation in the Universe can be understood by the cumulative emission from active and inactive galaxies in the Universe; there are no clear signs of additional source of the extragalactic background radiation. A large fraction of the background radiation can be resolved into individual sources.

9.7 Gamma-ray bursts

Discovery and phenomenology. In 1967, surveillance satellites for the monitoring of nuclear test ban treaties discovered γ-flashes similar to those that are expected from nuclear explosions. However, these satellites found that the flashes were not emitted from Earth but from the opposite direction—hence, these γ-flashes must be a phenomenon of cosmic origin. Since the satellite missions were classified, the results were not published until 1973. The sources were named gamma-ray bursts (GRB).
The flashes are of very different duration, from a few milliseconds up to ∼ 100 s, and they differ strongly in their respective light curves (see Fig. 9.58). They are observed in an energy range from ∼ 100 keV up to several MeV, sometimes to even higher energies.
The nature of GRBs had been completely unclear initially, because the accuracy with which the location of the bursts was determined by the satellites was totally insufficient to allow an identification of any corresponding optical or X-ray source. The angular resolution of these γ-detectors was many degrees (for some, a 2π solid angle). A more precise position was determined from the time of arrival of the bursts at the location of several satellites, but the error box was still too large to search for counterparts of the source in other spectral ranges.
Early models. The model favored for a long time included accretion phenomena on neutron stars in our Galaxy. If their distance was D ∼ 100 pc, the corresponding luminosity would be about L ∼ 1038 erg∕s, thus about the Eddington luminosity of a neutron star. Furthermore, indications of absorption lines in GRBs at about 40 and 80 keV were found, which were interpreted as cyclotron absorption corresponding to a magnetic field of ∼ 1012 Gauss—again, a characteristic value for the magnetic field of neutron stars. Hence, most researchers before the early 1990s thought that GRBs occur in our immediate Galactic neighborhood.
/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig58_HTML.jpg
Fig. 9.58
Gamma-ray light curves of various gamma-ray bursts; the different time-scales on the x-axis should be particularly noted. All these light curves appear to be very dissimilar. Credit: J.T. Bonnell, GLAST Science Support Center, NASA/Goddard Space Flight Center, Greenbelt, Maryland, USA
The extragalactic origin of GRBs. A fundamental breakthrough was then achieved with the BATSE experiment on-board the Compton Gamma Ray Observatory, which detected GRBs at a rate of about one per day over a period of 9 years. The statistics of these GRBs shows that GRBs are isotropically distributed on the sky (see Fig. 9.59), and that the flux distribution N( > S) clearly deviates, at low fluxes, from the S −1. 5-law. These two results meant an end to those models that had linked GRBs to neutron stars in our Milky Way, which becomes clear from the following argument.
Neutron stars are concentrated towards the disk of the Galaxy, hence the distribution of GRBs should feature a clear anisotropy—except for the case that the typical distance of the sources is very small ( ≲ 100 pc), much smaller than the scale-height of the disk. In the latter case, the distribution might possibly be isotropic, but the flux distribution would necessarily have to follow the Euclidean law $$N(> S) \propto S^{-3/2}$$ , as expected for a homogeneous distribution of sources, which was discussed in Sect. 4.​1.​2. Because this is clearly not the case, a different distribution of sources is required, hence also a different kind of source.
The only way to obtain an isotropic distribution for sources which are typically more distant than the disk scale-height is to assume sources at distances considerably larger than the distance to the Virgo cluster, hence D ≫ 20 Mpc; otherwise, one would observe an overdensity in this direction. In addition, the deviation from the  $$N(> S) \propto S^{-3/2}$$ -law means that we observe sources up to the edge of the distribution (or, more precisely, that the curvature of spacetime, or the cosmic evolution of the source population, induces deviations from the Euclidean counts), so that the typical distance of GRBs should correspond to an appreciable redshift. This implies that the total energy in a burst has to be E ∼ 1051 to 1054 erg. This energy corresponds to the rest mass Mc 2 of a star. The major part of this energy is emitted within ∼ 1 s, so that GRBs are, during this short time-span, more luminous than all other γ-sources in the Universe put together.
/epubstore/S/P-Schneider/Extragalactic-Astronomy-And-Cosmology/OEBPS/A129044_2_En_9_Fig59_HTML.jpg
Fig. 9.59
Distribution of gamma-ray bursts on the sphere as observed by BATSE, an instrument on-board the CGRO-satellite, during the about 9 year mission; in total, 2704 GRBs are displayed. The color of the symbols represents the observed strength (fluence, or energy per unit area) of the bursts. One can see that the distribution on the sky is isotropic to a high degree. Credit: G. Fishman et al., BATSE, CGRO, NASA
We note that the estimated energy of a GRB assumes that the relation between observed flux and luminosity is given by L = 4π D L 2 S. This relation is valid only for source which emit isotropically. We have seen that this assumption breaks down for some classes of objects, for example blazars, for which relativistic beaming plays a major role.
Identification and afterglows. In February 1997, the first identification of a GRB in another wavelength band was accomplished by the X-ray satellite Beppo-SAX. Within a few hours of the burst, Beppo-SAX observed the field within the GRB error box and discovered a transient source, by which the positional uncertainty was decreased to a few arcminutes. In optical observations of this field, a transient source was then detected as well, very accurately defining the position of this GRB. The optical source was identified with a faint galaxy. Optical spectroscopy of the source revealed the presence of absorption features at redshift z = 0. 835; hence, this GRB must have a redshift equal or larger than this. For the first time, the extragalactic nature of GRBs was established directly. In fast progression, other GRBs could be identified with a transient optical source, and some of them show transient radiation also at other wavelengths, from the radio band up to X-rays. The lower-energy radiation of a GRB after the actual burst in gamma-rays is called an afterglow.
With the launch of the SWIFT satellite in November 2004, the observations of GRBs entered a new phase. This satellite is equipped with three instruments: a wide-field gamma-ray telescope to discover the GRBs, an X-ray telescope, and an optical/UV telescope. Within a few seconds of the discovery of a GRB, the satellite targets the location of the burst, so that it can be observed by the latter two telescopes, obtaining an accurate position. This information is then immediately transmitted to the ground, where other telescopes can follow the afterglow emission and obtain spectroscopic information. In its first 8 years of operation, SWIFT discovered some 700 GRBs, of which ∼ 200 have their redshift determined. The afterglow could be studied in a much more homogeneous way than before. The prompt γ-ray emission carries about the same amount of total energy as the afterglow emission in the X-ray regime, whereas the total energy of the optical afterglow is smaller by a factor ∼ 100.
Relativistic motion. A GRB detected in May 1997 showed an afterglow also at radio frequencies. In the first ∼ 20 days, its radio light curve varied erratically, before it settled into a smoother behavior, with flux declining in time. The flux fluctuations in the initial light curve were interpreted as being due to scintillation in the inhomogeneous interstellar medium, very much like the scintillation in the Earth atmosphere.10 The end of the fluctuating period is then interpreted as being due to the growing size of the emitting source: Just like planets are not scintillating due to their large angular size on the sky, interstellar scintillations are observable only for sufficiently small sources. Hence, these observations provided a clear evidence for an expanding source responsible for the radio afterglow, as well as an estimate of the source size at the end of the scintillation period, of order a few light-weeks. Hence, the expansion velocity of the source must be of order of the speed of light.
There is another independent argument for the presence of relativistic motion in GRBs. The short time-scale of the γ-ray emission, together with its large flux, implies that the density of γ-ray photons in the source must be extremely high. In such a situation, the γ-rays are subject to a large opacity for  $$\mathrm{e}^{+}\mathrm{e}^{-}$$ -pair production—in other word, the γ-rays cannot escape from the source region, but are efficiently transformed into pairs. To escape this conclusion, Doppler boosting needs to be employed (cf. Sect. 5.​5.2). Allowing for relativistic velocities along our line-of-sight, the radiation density in the source declines significantly. Furthermore, the estimated source size, based on variability argument, increases if Doppler boosting is at work. With a Lorentz factor of the bulk motion of  $$\gamma = [1 - (v/c)^{2}]^{-1/2} \sim 10^{2}$$ or larger, the pair-production opacity constraints can be avoided, and source sizes of order 1013 cm can be accomplished.
Fireball model. Hence, GRBs are associated with a relativistic phenomenon, but the question of their nature still remained unanswered. One model of GRBs quite accurately describes their emission characteristics, including the afterglow. In this fireball model, the radiation is released in the relativistic outflow of electron-positron pairs with a Lorentz-factor of γ ≥ 100. This radiation is not isotropic, but most likely concentrated in a rather narrow beam, resembling the jets in AGNs. In order to form such collimated outflows, one needs a strong energy source, and presumably strong rotation whose rotation axis defines the preferred directions into which the jets flow. To collimate the jets, the presence of magnetic fields are probably also required.
Short vs. long-duration bursts. GRBs can be broadly classified into short- and long-duration bursts, with a division at a duration of t burst ∼ 2 s. The spectral index of the short-duration bursts is considerably harder at γ-ray energies than that of long-duration bursts. Until 2005, only afterglows from long-duration bursts had been discovered. Long-duration bursts typically occur in galaxies at high redshift, with a mean of z ∼ 2. 5. Also GRBs with very high redshift were discovered, with at least three having redshifts z > 6. One GRB redshift of z = 8. 2 has been spectroscopically obtained, and there are indications that an even higher-redshift burst was observed. In one case, an optical burst was discovered about 30 seconds after the GRB, with the fantastic brightness of V ∼ 9, at a redshift of z = 1. 6. For a short period of time, this source was apparently more luminous than any quasar in the Universe. In March 2008, a GRB at z = 0. 937 occurred which has a peak optical brightness of m = 5. 7—i.e., this source was visible for a very short period to the naked eye (it is not known, though, whether anyone peeked at the right position of the sky at that moment). Thus, during or shortly after the burst at high energies, GRBs can also be very bright in the optical.
Counterparts of long-duration GRBs: Hypernovae. In April 1998, the positional error box of a GRB contained a supernova, hinting for a possible connection. This has been verified subsequently, by finding that the light-curve of some optical afterglows were described by the sum of a declining power law in time plus the light-curve of a luminous supernova. For a GRB in March 2003, the presence of a supernova in the spectrum of the optical afterglow was identified, proving the direct connection between SNe and GRBs. Since most of the GRBs are located at high redshifts, the corresponding SN cannot be identified for them, but for more nearby long-duration bursts, the association is clearly established.
Long-duration GRBs are located in star-forming regions of galaxies, and their redshift distribution is similar to that of the star-formation rate density in the Universe.11 This observation yields a close connection of the GRB phenomenon to star formation, and thus the associated supernovae are due to young massive stars. Not every core-collapse SN yields a GRB, though. The current picture is that GRBs are produced in the core-collapse process of very massive stars, giving rise to extraordinarily energetic explosions, so-called hypernovae. The combination of stellar rotation and an internal magnetic field can form a highly relativistic bi-directional outflow after the collapse event, when the stellar material falls onto the newly formed compact remnant, a black hole. Even if the emission is highly anisotropic, as expected from the fireball model, the corresponding energy released by the hypernovae is very large.
Counterparts of short-duration GRBs. SWIFT has allowed the identification of afterglow emission from short-duration GRBs. In contrast to the long-duration bursts, some of these seem to be associated with elliptical galaxies; this essentially precludes any association with (core-collapse) supernova explosions. In fact, for one of these short burst, very sensitive limits on the optical brightness explicitly rules out any contribution from a supernova explosion. Furthermore, the host galaxies of short bursts are at substantially lower redshift, z ≲ 0. 5. Given that both kinds of GRBs have about the same observed flux (or energy), this implies that short-duration bursts are less energetic than long-duration ones, by approximately two orders of magnitude. All of these facts clearly indicate that short- and long-duration GRBs are due to different populations of sources. The lower energies of short bursts and their occurrence in early-type galaxies with old stellar populations are consistent with them being due to the merging of compact objects, either two neutron stars, or a neutron star and a black hole.
Footnotes
1
Readers not familiar with the optical/near-IR filter system may find it useful to consult Sect. A.4.2 in the Appendix at this point. We will also follow the usual practice and write R = 22 instead of R = 22 mag in the following.
 
2
Whereas the symbols for redshift and the z-band magnitudes are identical, we trust that no confusion will arise by that, as the meaning will always be clear by the context.
 
3
Note that a factor of 5 in magnification corresponds to a factor 25 in the exposure time required for spectroscopy. This factor of 25 makes the difference between an observation that is feasible and one that is not. Whereas the proposal for a spectroscopic observation of 3 h exposure time at an 8-m telescope may be successful, a similar proposal of 75 h would be hopelessly doomed to failure.
 
4
At the time of writing, ∼ 90 QSOs with z > 5. 7 and known, of which ∼ 40 have z > 6. 0; those were found from several wide-field imaging surveys, including SDSS, the CFHT quasar survey, the UKIRT Infrared Deep Sky Survey (UKIDSS), and Pan-STARRS.
 
5
The accuracy with which the position of a compact source can be determined is approximately given by the ratio of the FWHM and the signal-to-noise ratio with which this source is observed.
 
6
The reason for this bias is the very different K-correction in the sub-mm and radio regimes, due to the very different slopes of the spectral energy distribution in these two regimes, as can be seen in Fig. 9.30: Whereas the flux in the sum-mm regime increases as a source is moved to higher redshifts, its radio flux decreases strongly, thus biasing against the detection of high-z SMGs in the radio.
 
7
There a few TeV blazars at higher redshift, but as we discussed in Sect. 5.​2.​6, the featureless spectrum of most blazars renders the determination of a secure redshift sometimes uncertain.
 
8
We recall that the roughly equal energy in the optical and FIR extragalactic background radiation shows that about half of the cosmic star formation occurs in dust-obscured regions.
 
9
The derivation of the star-formation rate as a function of redshift is largely drawn from galaxy surveys which are based on color selection, such as LBGs, EROs and sub-mm galaxies. The possibility cannot be excluded that additional populations of galaxies which are luminous but do not satisfy any of these photometric selection criteria are present at high redshift. Such galaxies can be searched for by spectroscopic surveys, extending to very faint magnitude limits. This opportunity now arises as several of the 10-m class telescopes are now equipped with high multiplex spectrographs which can thus take spectra of many objects at the same time. One of them is VIMOS at the VLT, another is DEIMOS on Keck. With both instruments, extensive spectroscopic surveys are being carried out on flux-limited samples of galaxies. Among the first results of these surveys is the finding that there are indeed more bright galaxies at redshift z ∼ 3 than previously found, by about a factor of 2, leading to a corresponding correction of the star-formation rate at high redshifts. In a color-color diagram, these galaxies are preferentially located just outside the selection box for LBGs (see Fig. 9.4). Given that this selection box was chosen such as to yield a high reliability of the selected candidates, it is not very surprising that a non-negligible fraction of galaxies lying outside, but near to it are galaxies at high redshift with similar properties.
 
10
Recall that atmospheric scintillations are due to a space and time dependent refractive index of the air. For propagating radio waves, the same is true, except that the refractive index here is determined by the electron density of the ionized plasma in the ISM.
 
11
Indeed, it seems that the distribution of GRBs extends further out in redshift than that of the star formation density. This observational fact is most likely related to the finding that GRBs are found in host galaxies with small metallicity. It is possible that the metal enrichment of galaxies suppresses GRBs at later redshifts. The connection to the metallicity may have its origin on a possible metallicity-dependent star formation, i.e., allowing for higher-mass stars from metal-poor gas.